Human-Induced Changes in the Environment and Landscape of the Maltese Islands from the Neolithic to the 15th Century AD: as inferred from a scientific study of sediments from Marsa, Malta 9781407301204, 9781407331607

An in-depth study of man's impact on the environment and landscape of the Maltese islands from the Neolithic to Med

227 31 11MB

English Pages [167] Year 2007

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Human-Induced Changes in the Environment and Landscape of the Maltese Islands from the Neolithic to the 15th Century AD: as inferred from a scientific study of sediments from Marsa, Malta
 9781407301204, 9781407331607

Table of contents :
Front Cover
Title Page
Copyright
Table of Contents
List of Figures
List of Tables
ACKNOWLEDGEMENTS
Chapter I INTRODUCTION
Chapter II ENVIRONMENTAL BACKGROUND
Chapter III ARCHAEOLOGICAL BACKGROUND
Chapter IV STRATIGRAPHICAL INVESTIGATIONS
Chapter V RADIOCARBON DATING OF MARSA CORE 1
Chapter VI SEDIMENTOLOGICAL INVESTIGATIONS
Chapter VII BIOSTRATIGRAPHY AND PALAEOECOLOGY
Chapter VIII SYNTHESIS
APPENDIX I Munsell soil colours of the moist and dried samples of Marsa Core 1
APPENDIX II Shells found in Marsa Core 1, standardised per 150g sample
APPENDIX III Plant macro-remains of Marsa Core 1, not standardised
REFERENCES

Citation preview

Human-Induced Changes in the Environment and Landscape of the Maltese Islands from the Neolithic to the 15th Century AD as inferred from a scientific study of sediment from Marsa, Malta

Katrin Fenech

BAR International Series 1682 2007

Human-Induced Changes in the Environment and Landscape of the Maltese Islands from the Neolithic to the 15th Century AD as inferred from a scientific study of sediments from Marsa, Malta

Katrin Fenech

BAR International Series 1682 2007

B B A

b ........ h s l gu

i

g

BAR

PUBLISHING

his b

is

il bl

h B i ish ib

Contents List of figures List of tables Acknowledgements

iv v vii

Chapter I: Introduction 1.1. General 1.2. Location of the site

1 3

Chapter II: Environmental Background 2.1. Topography 2.1.1. Tertiary Geology 2.2. Tectonics – the structural setting of the Maltese Islands in the Strait of Sicily 2.2.1. Sub-recent tectonic movements 2.2.2. Historical earthquakes 2.2.3. Tectonics in the Marsa area 2.3. Climate 2.3.1. Climate and weather effects in the Marsa Plain in Modern Times 2.3.2. Factors inducing climatic change 2.3.3. Past climates of the Maltese Islands 2.3.4. Climate from ca 45,000 BP to the Late Glacial at around 15,000 BP 2.3.5. Climate from the Late Glacial at around 14,500 BP to the onset of the Holocene (around 11,400 BP) 2.3.6. Early Holocene 2.3.7. Late Holocene (around 4500 BP to present), central Mediterranean and Malta 2.4. Geomorphology 2.4.1. General 2.4.2. Rock erosion 2.4.3. Karst features 2.4.4. Soils 2.4.5. Soil erosion 2.5. Hydrology 2.5.1. Surface water – fluvial drainage 2.5.2. Surface water in the Marsa catchment 2.5.3. Subsurface water 2.5.4. Underground water resources in the Marsa catchment 2.6. Sea level and coastal changes since the Pleistocene 2.6.1. Global (eustatic) changes 2.6.2. Isostatic changes in the sea level Chapter III: Archaeological Background 3.1. Early descriptions of monuments and history of archaeological research in the Maltese Islands 3.2. Chronology 3.2.1. Radiocarbon dating 3.2.2. The Maltese radiocarbon dates 3.3. Chronology, cultural sequence and settlement development in the Maltese Islands 3.3.1. The Neolithic Period 3.3.1.1. Ghar Dalam Phase (5500-4100 2σ cal. BC) 3.3.1.2. Grey Skorba Phase (traditionally dated 4500-4400 BC) 3.3.1.3. Red Skorba Phase (4350-3650 2σ cal. BC) 3.3.2. The Temple Period 3.3.2.1. Zebbug Phase (4350-3050 2σ cal. BC) 3.3.2.2. Mgarr Phase (3700-2900 2σ cal. BC) 3.3.2.3. Ggantija Phase (3360-2940 2σ cal. BC) i

6 6 8 10 10 12 12 12 13 13 14 15 16 17 18 18 20 21 21 22 23 23 25 26 28 28 28 30

32 33 33 34 35 37 37 38 39 39 39 40 41

3.3.2.4. Saflieni Phase (traditionally placed around 3300-3000 BC) 3.3.2.5. Tarxien Phase (3150-2450 2σ cal. BC) 3.3.3. The Bronze Age 3.3.3.1. Tarxien Cemetery Phase (2900-1420 2σ cal. BC, Castelluccio culture in Sicily, 3080-1420 2σ cal. BC) 3.3.3.2. Borg in-Nadur Phase (ca. 1500-700 BC, Thapsos culture in Sicily 1880-830 2σ cal. BC) 3.3.3.3. Bahrija Phase (ca. 900-700 BC) 3.3.4. The Historical Period 3.3.4.1. The Phoenician/ Punic Period (8th century BC – 218 BC) 3.3.4.2. The Roman Period (218 BC – 535 AD) 3.3.4.3. The Arab Period (870-1091 AD) 3.3.4.4. Post-Arab Period (1091-1282 AD) 3.3.4.5. Spanish Rule (1282-1530 AD) 3.3.4.6. Early Modern Malta – the Knights Period (1530-1798 AD) 3.3.4.7. French Period (1798-1800 AD) 3.3.4.8. British Period (1800-1964 AD) 3.3.4.9. Independent Malta 3.4. Population estimates 3.5. Agriculture, local resources and economic development from the Neolithic up to the 15th century AD 3.5.1. The Neolithic Period 3.5.2. The Temple Period 3.5.3. The Bronze Age Period 3.5.4. The Phoenician/ Punic Period 3.5.5. The Roman Period 3.5.6. The Arab and Post-Arab Period 3.5.7. Spanish Rule 3.6. Changes and developments in the Marsa area since Early Modern Malta 3.6.1. The French Period 3.6.2. From the British Period to present day

41 42 42 42 43 44 45 45 46 48 49 49 50 50 50 51 51 52 52 53 54 54 55 57 58 59 60 60

Chapter IV: Stratigraphical Investigations 4.1. Introduction 4.2. General stratigraphy inferred from borehole surveys 4.3. Detailed description of the stratigraphy of Marsa Core 1 and Marsa Core 2 from the Marsa Sports Ground 4.4. Description of the Harrison & Company Cores, Marsa Sports Ground 116 4.5. Results 4.6. Interpretation 4.7. Conclusion

65 66 67 67 68

Chapter V: Radiocarbon dating of Marsa Core 1 5.1. Introduction 5.2. Results 5.3. Discussion

69 69 69

Chapter VI: Sedimentological Investigations 6.1. Introduction 6.2. History of sediment studies in the Maltese Islands 6.3. Particle size distribution 6.3.1. Methods 6.3.2. Water content 6.3.3. Results 6.4. Organic content 6.5. X-ray fluorescence (XRF) Analysis – major and minor elements by quantity ii

62 62

72 72 73 73 73 74 75 75

6.5.1. Results 6.6. Magnetic susceptibility 6.6.1. Material and methods 6.6.2. Results 6.7. Discussion

76 78 79 79 80

Chapter VII: Biostratigraphy and Palaeoecology 7.1. Molluscs 7.1.1. Introduction 7.1.2. Setting the scene 7.1.2.1. Data from Quaternary deposits 7.1.2.2. Data from archaeological excavations until 1971 7.1.2.3. Quantitative data from environmental samples in archaeological contexts 7.1.3. The environmental background 7.1.4. Preservation of shells found in Marsa Core 1 7.1.4.1. Fracturing 7.1.4.2. Blackening 7.1.4.3. Erosion and weathering of the shells 7.1.5. Notes on identification 7.1.6. Nature of the molluscan assemblages 7.1.7 Grouping of the shells 7.1.7.1. By broad habitat categories and salinity tolerance 7.1.7.2. Land snails according to their habitat requirements 7.1.7.3. Brackish water and marine molluscs according to their benthic habitats 7.1.7.4. Burrowing land snails 7.1.7.5. Number of species 7.1.7.6. Statistical analyses 7.1.8. Analysis of faunal change throughout the core 7.1.9. Notes on selected species 7.2. Vertebrates 7.2.1. Introduction 7.2.2. Identification and results 7.2.3. Conclusion 7.3. Palaeobotany 7.3.1. Introduction 7.3.2. Pollen Analysis 7.3.2.1. Material and methods 7.3.2.2. Results 7.3.3. Plant macro-remains 7.3.3.1. Results 7.3.4. Analysis of vegetational change throughout the core 7.4. Charcoal – black carbon 7.4.1. Introduction 7.4.2. General considerations 7.4.3. Limitations of the interpretation 7.4.4. Natural and anthropogenic fires 7.4.5. Material and methods 7.4.6. Results

81 81 81 81 82 84 84 85 85 86 86 86 86 87 87 88 89 90 90 91 92 93 93 93 93 94 94 94 95 95 96 97 98 98 99 99 99 100 101 101 102

Chapter VIII: Synthesis 8.1.Introduction 8.2. Section 1: from the Late Pleistocene to the Neolithic 8.3. Section 2: From the Neolithic to Byzantine Malta 8.4. Section 3: From Byzantine Malta to the 15th century AD 8.5. Conclusion

104 104 107 111 113

Appendix I: Munsell soil colours of Marsa Core 1 Appendix II: Shells of Marsa Core 1, standardised per 150g sample Appendix III: Plant macro-remains of Marsa Core 1, not standardised References

115 120 133 139

iii

List of Figures Figure 1.1.: Figure 1.2.: Figure 1.3.: Figure 1.4.1.: Figure 1.4.2.: Figure 1.4.3.: Figure 1.5.: Figure 2.1.: Figure 2.2.: Figure 2.3.: Figure 2.4.: Figure 2.5.: Figure 2.6.: Figure 2.7.: Figure 2.8.: Figure 2.9.: Figure 2.10.: Figure 2.11.: Figure 2.12.: Figure 2.13. Figure 3.1.: Figure 3.2a.: Figure 3.2b.: Figure 3.2c.: Figure 3.3a.: Figure 3.3b.: Figure 3.3c.: Figure 3.3d.: Figure 3.4a.: Figure 3.4b.: Figure 3.4c.: Figure 3.5a.: Figure 3.5b.: Figure 3.6.: Figure 3.7a.: Figure 3.7b.: Figure 3.8a.: Figure 3.8b: Figure 3.8c: Figure 3.9.: Figure 4.1.: Figure 4.2.: Figure 4.3.:

Location of the Maltese Islands in the central Mediterranean Sea. After Schembri, 1993. Landforms of Malta with some localities. After Nehring, 1966. Topographic map of the area around the Marsa Sports Ground, showing the location of the boreholes. After MEPA, 2003. Drawing of section exposed by excavation works of Ballut Blocks. Photo of section exposed by Ballut Blocks facing the car park. Photo of section located ca. 5m further west from Fig. 1.4.1. Map of Gozo with some localities mentioned in the text. Geology of Malta with Marsa catchment outlined. After House et al., 1961. The Maltese Islands in relation to the present geodynamic framework. Source: Pedley et al., 2002. Fault pattern of the Maltese Islands with the Marsa catchment outlined. After Reuther, 1984. Comparison between GISP 2 record and palaeoenvironmental reconstructions from Ocean Drilling Program (ODP) Site 976. Source: Combourieu Nebout et al., 2002. Location of the various sites with palaeoclimatic data discussed in the text. Down-core records of different sea surface temperature (SST) proxies in core AD91-17 from the Southern Adriatic Sea and corresponding climatic events. After Sangiorgi et al, 2003b. Steepness of slopes in Malta with Marsa catchment outlined and some localities. After Bowen-Jones et al., 1961. Extract from a map of Malta by Jean Quintin (Quintinus) published in 1536 showing the Marsa Hortus and two freshwater tributaries. From Agius-Vadala & Ganado, 1956. Drainage patterns and valley systems of the Maltese Islands and their extensions below present day sea level. Source: Vossmerbäumer, 1972. Physiography of Malta with main river valleys and Marsa catchment outlined. Topographical map showing a detailed drainage pattern of the Marsa catchment with floodplain. Hydrology of Malta after T.O. Morris, 1952, with Marsa catchment and Marsa Sports Ground outlined. Oxygen isotope record for the past 2.6 million years deduced from benthonic foraminifera of the ODP core 677 with labels of selected isotope stages added for orientation. Odd numbers correspond to interglacials, even numbers to glacial stages. Source: Pirazzoli, 1998. Chronological timescale of the different cultural phases in the Maltese Islands from the Neolithic to Early Modern Malta (Knights Period). Settlement development in Malta during the Ghar Dalam Phase. Settlement development in Malta during the Grey Skorba Phase. Settlement development in Malta during the Red Skorba Phase. Settlement development in Malta during the Zebbug Phase. Settlement development in Malta during the Mgarr Phase. Settlement development in Malta during the Ggantija Phase. Settlement development in Malta during the Tarxien Phase. Settlement development in Malta during the Tarxien Cemetery Phase. Settlement development in Malta during the Borg in-Nadur Phase. Settlement development in Malta during the Bahrija Phase. Settlement development in Malta during the Phoenician/PunicPeriod. Settlement development in Malta during the Roman Period. Topographic map of the Marsa area, showing the location of the boreholes and the various archaeological sites. Source of base map: MEPA. Settlement development in Malta during the Arab Period. Settlement development in Malta during the Spanish Period up to the Knights Period. Source: Wettinger, 1975. Settlement growth in Malta by 1842. Source: Richardson, 1961. Settlement growth in Malta by 1956. Source: Richardson, 1961. Settlement growth in Malta by 2004. Source: EEA, 2004. Amphorae occurrence during various phases of the Roman Period in Malta. Location of the different transects in the Marsa/ Grand Harbour Area and of the two cores (Marsa Core 1 and Marsa Core 2) taken during the present study. Graphic representation of the stratigraphy and correlation of Marsa Cores 1 and 2 and the three cores removed from the Marsa Sports Complex by Harrison & Co during geotechnical studies in connection with the construction of the Marsa Sports Complex. Graphic representation of Transect 1, showing the stratigraphy and correlation of five cores excavated by Malta Shipbuilding during geotechnical studies in connection with the building of the shipbuilding yard in 1976. iv

Figure 4.4.: Figure 4.5.: Figure 4.6.: Figure 4.7.: Figure 5.1.: Figure 5.2.: Figure 6.1.: Figure 6.2.: Figure 6.3.: Figure 6.4.: Figure 7.1.1.: Figure 7.1.2.: Figure 7.1.3.: Figure 7.1.4.: Figure 7.1.5.: Figure 7.1.6.: Figure 7.1.7.: Figure 7.1.8.: Figure 7.3.1.: Figure 7.3.2.: Figure 7.4.1.: Figure 8.1.: Figure 8.2.: Figure 8.3.:

Graphic representation of Transect 2, showing the stratigraphy and correlation of four cores excavated by Malta Shipbuilding during geotechnical studies in connection with the construction of the shipbuilding yard in 1976. Graphic representation of the stratigraphy and correlation of two cores removed from Il-Menqa at Marsa during geotechnical studies in connection with the construction of the new Maltacom building. Topographic map of the inlet at Marsa, into which the dock of Malta Shipbuilding had been constructed, showing the exact locations of the cores studied in Transect 1 and Transect 2. Part of Sample 170 of Marsa Core 1, at 980cm, showing lamination. Aerial photograph most likely taken in the early 1930s, showing the Marsa Sports Ground and race track in the foreground, and the Grand Harbour and inlet at Marsa behind. Depth-age relationships for calibrated (2 sigma) radiocarbon dates from Marsa Core 1, and radiocarbon results (determined by Beta Analytic Inc.) of MRS 1-007 and MRS 1-058. Water content, organic matter and particle size distribution throughout Marsa Core 1. XRF determined metal components in Marsa Core 1, expressed in μg/g for every fifth sample. XRF determinations of major elements in Marsa Core 1 for every fifth sample, amounts expressed in μg/g. Volume frequency dependent and mass specific magnetic susceptibility and some XRF determined elements (expressed in μg/g) for every fifth sample, Marsa Core 1. Variations of habitat specific assemblages throughout Marsa Core 1, expressed as percentages of the total number of shells. Variations of molluscan assemblages throughout Marsa Core 1 according to their broad habitat category, with euryhaline and polyhaline species grouped together as brackish water. Land snail distribution throughout Marsa Core 1 according to broad habitat category, expressed in numbers per 150g sample. Distribution of epifaunal, epi/infaunal and infaunal marine molluscs in Marsa Core 1, standardised per 150g/ sample and expressed as percentage values. Distribution of burrowing land snails belonging to the family of Ferussaciiadae throughout Marsa Core 1, standardised per 150g sample. Species richness and abundance in Marsa Core 1. Non-metric multidimensional scaling ordination of the molluscan assemblage from all samples from Marsa Core 1, standardised per 150g/sample, based on a similarity matrix constructed using the BrayCurtis similarity index. Non-metric mulitdimensional scaling ordination of the molluscan assemblages from Marsa Core 1 based on presence/ absence of broad habitat types (land, fresh water, brackish water, marine), in the samples. Marsa Core 1, Pollen Analysis (except herbaceous taxa), count expressed as a percentage of the total pollen sum. Marsa Core 1, Pollen analysis of the predominantly herbaceous taxa, count expressed as a percentage of the total pollen sum. Macroscopic and microscopic charcoal remains throughout Marsa Core 1. Extrapolated dates and sedimentation rates of Marsa Core 1 between 750cm and 320cm. Holocene rise of sea levels according to data from Sicily and nearby Italy, and transposed on Marsa Core 1. Marsa Core 1, pollen analysis of selected taxa, counts expressed in total number.

List of Tables Table 3.1.: Table 3.2.: Table 3.3.: Table 5.1.: Table 7.1.1.: Table 7.1.2.: Table 7.1.3.: Table 7.1.4.: Table 7.2.1.:

Chronological overview of the cultural sequences in Malta. Radiocarbon dates from Malta. Radiocarbon dates from Sicily from sites with pottery styles that find parallels in Malta during various prehistoric phases. Details of the AMS radiocarbon dates from Marsa Core 1. Land snails found in various Quaternary deposits of the Maltese Islands. Land snails from different archaeological contexts and deposits until 1971. Land snails from environmental samples of archaeological deposits. Molluscan species according to their salinity ranges. List of vertebrate remains found in Marsa Core 1.

v

ACKNOWLEDGEMENTS It has been a great privilege for me to have been jointly supervised by Prof. Anthony Bonanno of the Department of Classics and Archaeology, and by Prof. Patrick J. Schembri of the Department of Biology, of the University of Malta. I am most grateful to them, for their expertise and guidance ensured throughout that it was also a rewarding personal experience for me to write this thesis. Palaeo-environmental researchers from other universities also contributed by readily making available their published research results and many took the time to discuss various interpretations of the results. I am most grateful here to Dr. Christopher O. Hunt of Queens University, Belfast, who showed me the multifaceted possibilities of data interpretations. I am also very grateful to Prof. Anthony Frendo, who, when I was a first year student of Archaeology at the University of Malta, gave me the opportunity to specialise in Environmental Archaeology. He also introduced me to the principle of ‘serendipity’. Through this, I came across many important books and articles and often also their authors, who readily discussed various aspects of my work with me. Among these, I am particularly grateful to Prof. José Carrión of Murcia University (Spain), Prof. Claus-Dieter Reuther of Hamburg University, Dr. Willy Tinner of Berne University, Dr. Charles Turner of Cambridge University, and Dr. David Trump. Prof. Wighart von Königswalde at the Palaeontological Institute Bonn and Prof. Wolfgang Böhme of Museum Koenig kindly helped identifying several bone fragments. I am also grateful to Dr. Nicholas Vella of the Department of Classics and Archaeology, Malta, Dr. Reuben Grima of Heritage Malta and Dr. Anton Bugeja for providing me with helpful material and advice. Often, much needed help and encouragement also came from non-academic quarters. My husband Ivan, my daughter Kyra, my father and my sister Julie were an incredible source of support, as were my friends. This work forms part of a larger study on the palaeo-environment of the Maltese Islands and the role of humans in bringing about environmental change. This was being undertaken by the Department of Classics & Archaeology and the Department of Biology of the University of Malta and the Queen’s University of Belfast. I am again very grateful to Dr Christopher O. Hunt of Queen’s University, Belfast, for his invaluable advice on evaluating the potential of Marsa as a coring site for the study. The sediment cores, Marsa Core 1 in particular, also form literally the core of this study – I am therefore very grateful to Ms Linda Eneix, who through the OTS Foundation, sponsored the mechanical retrieval of the cores from Marsa. All material and equipment necessary for the processing of the cores was generously provided by my co-supervisor, Prof. Patrick J. Schembri of the Department of Biology at the University of Malta, while my other supervisor, Prof. Anthony Bonanno of the Department of Classics and Archaeology at the University of Malta, made many papers and books available for my research. The cores retrieved from Marsa were split into two halves and one half was given to Frank Carroll of Queen’s University, Belfast, for his own PhD research. I am most grateful to Frank Carroll, for not only did he kindly make available the data he generated so far, but also spent many hours discussing results and interpretations with me. All the data gathered from the Marsa Core 1 samples would have been very limited in their meaning and interpretation, had it not been for the procurement of crucial radiocarbon dates. These were generously sponsored by Ms Linda Eneix of the OTS Foundation, Dr. Christopher O. Hunt through the University of Huddersfield, Prof. Anthony Bonanno and Prof. Patrick J. Schembri through the University of Malta and, in his personal capacity, my father, Klaus Hämmerlein. Darden Hood and Ron Hatfield of Beta Analytic Inc. in Florida were very helpful in explaining the measured dates. I am very grateful to all the persons and for all the circumstances that allowed the research to develop the way it did. The following work is the result of these many complex interactions. It is hoped that it contributes positively to the current knowledge of the past environment of the Maltese Islands and perhaps starts new discussions that lead to further research on a wider scale.

vii

intensive tree felling through slash-and-burn to gain precious agricultural land for crops (Grech, 2001: 48). Timber was probably also used as a central column at Skorba in the construction of early Ghar Dalam mudbrick huts (Trump, 1966: 10) while the wood of Cercis siliquastrum (Judas tree), Crataegus sp. (Hawthorn) and Fraxinus sp. (Ash) was used for domestic fires (Metcalfe, 1966: 54). Human-induced fires, aimed to remove the spiny and distasteful undershrubs and to promote the growth of edible plants and young shrubs for sheep and goat (Grove & Rackham, 2001: 229) may also have been employed. The effect of browsing in prehistoric Malta still needs to be established, as this depends on the size of the herds and whether they were controlled or allowed to browse indiscriminately. Though sheep and goats seem to do little damage to fully grown trees once the trees have grown out of reach, young and edible trees fall victim and tree regeneration is inhibited (ibid.). Often, writers on the Temple Period connect the sudden collapse of the Temple Culture with a severe degradation of the environment due to over-exploitation of the resources and subsequent soil erosion (for example Evans, 1959: 39; Trump, 1966: 51; Grech, 2001: 84), possibly compounded by a succession of dry rainless years (Bonanno, 1986b: 40). Pollen analyses from a Bronze Age cistern at Tal-Mejtin in Luqa by Godwin (1961: 8) led to the conclusion that the Bronze Age environment then was ‘already’ very much like it is today: grass and low herbs, Mediterranean scrubs and very few trees (Trump, 2000: 99-100). A more recent analysis of molluscan remains from a Zebbug Phase tomb at the Xaghra Stone Circle revealed also an open country/steppe vegetation in this earliest phase of the Temple Period. (Malone et al., 1995; Schembri &

Chapter I INTRODUCTION 1.1. General The Maltese Islands occupy with a total surface of 316 km² a central place in the Mediterranean Sea at latitude 35º48’ – 36º05’ North and longitude 14º11’ – 14º35’ East of the equator (Vossmerbäumer, 1972: 4). The largest island is Malta (246 km²), followed by Gozo (67 km²) and Comino and several smaller, uninhabited islets (Azzopardi, 1995: 19-21). Located ca. 80 km south of Sicily and 290 km east of Tunisia, the Libyan coast is 355 km to the south, and the nearest form of land to the west is the island of Crete, at a distance of 855 km (see Figure 1.1, Trump, 2002: 15). The environment of the Maltese Islands is generally thought to have been extensively forested prior to the arrival of the first settlers in the Neolithic. Hunt’s 1996 palynological investigations of Quaternary deposits at Fiddien Valley in Malta shows a scrubby vegetation of Corylus, Pinus, Alnus, Ostrya, Filicales and Cyperace, with other broad-leaved tree and shrub species, but also patches with open-ground taxa during the Pleistocene interstadials. The Fiddien tufa deposit unfortunately lacks an absolute date, and could be of any age from Late Pliocene to Early Holocene (Hunt, 1997: 102-3), although it is most likely that the deposition took place during a mid to late Pleistocene interglacial phase (Pedley et al., 2002: 96). The introduction of agriculture by these first settlers is believed to have been accompanied by

Figure 1.1.: Location of the Maltese Islands in the central Mediterranean Sea. After Schembri, 1993.

1

Hunt, forthcoming). The result begs the question, whether the often mentioned tree cover could have fallen victim already in the Ghar Dalam Phase, the earliest period of the settlers. It emphasizes that the magnitude of the anthropogenic influence on the Maltese environment needs to be assessed with newly obtained data. These data would allow the reconstruction of the environment prior to the arrival of the first settlers and then to establish any changes it experienced, whether anthropogenic or through natural forces, or a combination of the two.

these sediment layers span over a considerable length of time, although the rate of deposition – and retention of deposition – also depends in part on topographical features. Scientific analyses of undisturbed lake deposits have resulted in very detailed reconstructions of the environment in many parts of the globe (e.g. Ramrath et al., 2000; Hallett et al., 2003; Sadori et al. 2004), but the absence of proper lacustrine bodies renders this method unsuitable in a Maltese context. Marine cores are also suitable (e.g. Harle, 1997; Pitman & Ryan, 2001), yet despite the availability of suitable deposits in the Maltese Islands, here financial constraints due to the advanced technical equipment needed for the retrieval of the cores from the seabed put this method out of the reach of the present study.1 Important data and results can also be gained from the study of stratigraphic sequences in

A detailed and scientific study of anthropogenic and other influences that shaped the Maltese environment during the Holocene has to rely on the uncontaminated retrieval of undisturbed sediments. The longer this vertical stratigraphic sequence is, the higher are the chances that

Figure 1.2.: Landforms of Malta with some localities. After Nehring, 1966.

1 A quotation from Bezzina & Sons, Marsa, for the retrieval of a core from the deepest trough in the Grand Harbour, to be taken from a barge, amounted to LM 5000 in 2002, despite generous discounts from the contractor.

2

exposed sections, as for example from studies of stratigraphic sequences dated to the Upper Pleistocene and Holocene in Lampedusa (Giraudi, 2004) and Tunisia (Zielhofer et al., 2004). Alluvial environments in particular have the potential to preserve important data on changes in natural environments in their sediments. Here, the Mediterranean record is particularly suitable, as the climate makes the environment highly susceptible to erosion as a result of a combination of infrequent but intense rainstorms and a vegetation cover that is limited by drought, grazing and fire (Brown, 1997: 237). Through the resulting hillwash and subsequent deposition of the eroded soil, a depositional sequence can be traced. In Malta, there are several alluvial plains that have the potential to preserve a sedimentary record, among which are two extensive flood plains, namely the Burmarrad Plain near Salina Bay and the Marsa Plain that adjoins the Grand Harbour (Figure 1.2.). However, the level of preservation and the sedimentary time scale depend on the depth of the underlying bedrock and the surrounding topography.

needs to be assessed on the basis of scientific data in addition to conventional means. Through the scientific study of sediments and their components, valuable additional information about human interactions with the Maltese environment and their responses to natural and anthropogenic changes may be gained. It is suggested that this may perhaps result in a greater understanding of the developments and changes, detectable through the material remains, of past Maltese cultural periods and phases. The present study thus aims to reconstruct the Holocene environment of the Maltese Islands through the scientific study of sediments from Marsa, Malta and highlight the contributions such an inter-disciplinary study can make to archaeology. 1.2. Loction of the site A very good potential record for the present study was suspected to be preserved in the largest alluvial plain of the Maltese Islands in an area referred to as Marsa. This flood plain stands out among the other flood plains in Malta for a unique combination of geographical and archaeological factors. It has the largest water catchment and is surrounded by numerous archaeological sites that testify anthropogenic presence at least from the Ggantija Phase onwards (traditionally placed at 3600-3000 BC, see below). When an application to demolish an existing building and to construct new premises on two floors at the Marsa Sports Ground (located in the floodplain of

The ‘classical’ approach to research on past Maltese cultures has been based mainly on the study of material remains such as pottery and architectural features (e.g. Trump, 1966; Evans, 1971; Trump, 2002). However, past societies function – as we do today – within their environment. They interacted, acted and reacted with it. The extent of these interactions, actions and reactions

Figure 1.3.: Topographic map of the area around the Marsa Sports Ground, showing the location of the boreholes. Of the three cores retrieved from here, Marsa Core 1 forms the basis of the present study. Altitude contours every 2.5 m. After MEPA (Malta Environment and Planning Authority), 2003

3

Marsa) was filed at the Malta Environment and Planning Authority (MEPA) in 2000,2 preliminary investigations into the depth of the sediments revealed that bedrock was found at around 11.5m below ground level at the chosen construction site.3

levelled. Then, these holes were filled with concrete and the site was left abandoned. A quick investigation of the scattered sediments in April 2002 revealed the potential of the site: apart from an array of different marine and terrestrial soils, often littered with shells but sometimes also appearing sterile or just stony, the majority of the sediments had furthermore been waterlogged due to the low topography of the site.6 Waterlogging is important for the preservation of pollen grains, which rapidly decay in alkaline soils under aerobic conditions. Human activity in the area could be ascertained through the presence of several pottery sherds, the earliest dating back to the Ggantija phase, ca. 3600-3000 BC.7 After showing the site to Dr Chris Hunt8 and Frank Carroll,9 it was decided to take a test core to establish scientifically the suitability of the sediments. The core was extracted with a hand soil auger, designed by Eijkelkamp of the Netherlands. It was possible to penetrate the sediments to a maximum depth of 3.52m, which is equivalent to 1.72m below present day sea level. Beyond that depth, the sandy-clayey sediments filled the borehole as soon as the core was extracted, thus preventing further penetration. Biological investigation of this 3.52m core showed that the sediments underlying those levelled by bulldozers were undisturbed. It also showed that any core suitable for the biological and

The site in question lies at the eastern end of Wied ilGonna and north of an area known as Ta’ Ceppuna (Figure 1.3.). The MEPA case officer’s report also highlights the deep sedimentation,4 citing the Draft Grand Harbour Local Plan (MEPA, 1997), where the site is described as being an important geomorphological unit, also of scientific significance in hydrology, geomorphology and Quaternary palaeontology.5 The few buildings found in the vicinity of this spot are not constructed on bedrock (and thus are only single storey buildings), but rest, like the adjacent street, on a layer of clinker of varying thickness. However, as the new building was planned to be much larger and on two floors, a large number of piles were necessary. After the excavation and removal of the uppermost 1.5m, which consisted of concrete, limestone chippings and clinker (Figure 1.4.1.-3.), these piles were constructed by Ballut Blocks Ltd and varied in depth between 12 and 14 meters. The sediments that had been excavated for the construction of the holes for these piles were scattered around the various holes and the ground was afterwards E ÅÆW

E ÅÆW

Figure 1.4.1.: Drawing of section exposed by excavation works of Ballut Blocks. This section was at 1.08m distance from the borehole location of Marsa Core 1 (see Figure 1.3.).

Figure 1.4.3.: Photo of section located ca. 5m further west from Fig. 1.4.1. (see Figure 1.3.), highlighting the variations in the slag and limestone deposits. The underlying alluvial deposit appears homogenous. Height from ground level to black cable underneath the stone, ca. 1.20m.

Figure 1.4.2.: Photo of section exposed by Ballut Blocks facing the car park (see Fig. 1.3.). Top layer here is concrete, two layers of slag visible. On the left, concrete pile.

6 Before excavation and construction of the piles, during which roughly 1.20m of sediments were removed, the ground level was 2.50m above present day sea level. 7 The pottery sherds were identified by Nathaniel Cutajar, then Curator of the Museum of Archaeology, Valletta. All dates of phases and periods follow the table in Bonanno, 2000: 5. 8 Senior Lecturer, Dept. of Archaeology and Palaeoecology, Queens University Belfast, UK. 9 PhD research student at Queens University, Belfast, UK.

2

Development Application 000930/00, dated 23rd February, 2000. The site investigation was done by Harrison & Co (Malta) Ltd. 4 A.Borg, 2001; 2. 5 These observations were subsequently also published on p. 66 of the final version of the Grand Harbour Local Plan (MEPA, 2002a). 3

4

chemical investigation of the sediments would have to be retrieved by means of a mechanical corer, which was able to prevent the sediments from collapsing by means of a sleeve that could be extended and pushed lower as necessity demanded. In June 2002, A.N. Terracore Ltd, a company experienced in geological site investigations, was commissioned to extract several columns of sediments from this site within the Marsa Sports Ground. They extracted three cores, henceforth named Marsa Core 1, 2 and 3 (see Figure 1.3. and below). Of these cores, the results obtained from Marsa Core 1 will form the basis of the following elaborations and discussion.

Figure 1.5.: Map of Gozo with some localities mentioned in the text. 1 = Ggantija Temples, 2 = Xaghra Circle, 3 = Sta. Verma, 4 = Xewkija, 5 = Xaghra, 6 = Xlendi, 7 = Marsalforn, 8 = Victoria, 9 = Wardija Point, 10 = Ta‘ Sarraflu, 11 = Qawra.

5

Pedley et al. (2002) published a popular geological outline of the Maltese Islands.

Chapter II ENVIRONMENTAL BACKGROUND

2.1.1. Tertiary geology 2.1. Topography The Maltese Islands consist entirely of Tertiary limestones with subsidiary clays and marls, all laid down beneath the sea like a layered cake during various stages of the Oligo-Miocene between around 30 to 5 million years ago (Pedley et al.: 2002: 41). Unlike in Pantelleria and Linosa as well as on the Ragusa Plateau in Sicily, there are no magmatic rocks (Vossmerbäumer, 1972: 11). In all, five different formations can be identified and their succession on the islands is as follows (after Reuther, 1984a: 68.) 5. Upper Coralline Limestone (up to at least 162 m thick) 4. Greensand (0-12 m thick) 3. Blue Clay (0-65 m thick) 2. Globigerina Limestone (25-209 m thick) 1. Lower Coralline Limestone (over 140 m, base not seen above sea level)

The present topography of the Maltese Islands is the result of a complex interplay between the geology, tectonic movements, hydrology, sea level variations, and erosion by wind, sun and water, driven by climate, weather and anthropogenic actions. To understand or evaluate any anthropogenic interventions and the impacts these interventions would have had on the environment of the Maltese Islands, it is vital to place the interventions within a dynamic frame provided by the natural environment. In this chapter, the various components of the environment will be discussed for the Maltese Islands in general and, where possible, for the Marsa area in particular. The possibly oldest systematic description of the topographical features of the Maltese Islands that goes beyond a traveller’s account dates back to 1647, when the historian Gian Frangisk Abela published his description of the Maltese Islands. In 1783, the French Knight of the Order of St. John and well-known geologist, Déodat de Dolomieu published a more scientific account of the temperature and climate in Malta as part of a description of the Aeolian Islands. In 1854, Admiral T.A.B. Spratt published the first geological descriptions of the Maltese Islands in the form of a pamphlet, soon followed by the first geological maps by Earl of Ducie in the same year and by A.L. Adams in 1870 (Hyde, 1955: 5). Gulia (1858-59) also discussed Maltese geology, as did several other authors, but these accounts mainly focussed on the more sensational ossiferous fissures and Quaternary deposits (e.g. Adams, 1863, 1866, 1879; Hutton, 1866 and later Cooke, 1891 and Trechmann, 1938). Fuchs (1875) discussed in detail the Tertiary geology, but possibly due to the German language, only his list of fossils was widely used (Hyde, 1955: 6). As a member of the “Challenger Expedition”, Sir John Murray published one of the most important and detailed works on Maltese geology in 1890. The publication was so thorough that by the end of the 19th century, many important geological facts were known (ibid.). In the 20th century, research on the general geology again became particularly intense after World War II, when Hyde (1955) re-investigated the water supply and the possibility of oil exploitation. A concise and clear account of the geography was provided by House et al. (1961), and this still forms the backbone of many recent environmental studies (e.g. Planning Authority, 1990). Vossmerbäumer (1972) made a very thorough study of the geology, but here again the German language may have acted as a restriction to limit citation of this work in more recent publications. Recent studies of the tectonics and faulting by Reuther (1984a, 1984b, 1987) Grasso & Reuther (1988), and Reuther et al. (2002) contributed more information on the past formation of the islands, while Pedley published over twenty technical papers on Maltese geology between 1974 and 2000 (see references in Pedley et al., 2002: 104-5). Very recently,

This succession bears a strong resemblance to that of south-east Sicily, to which Malta was connected via a land bridge during the last glacial period, and to the coastal area of Tunisia in North Africa (House et al., 1961: 26). As all these formations are also present within the Marsa catchment (see Figure 2.1.), they will be briefly described: 1. The Lower Coralline Limestone is a hard, pale grey limestone unit. It forms cliffs up to over 100 m high, especially along the western and south-western coast of the islands (see Figure 2.1) and is the oldest exposed rock formation on the Maltese Islands (Pedley et al., 2002: 35, 43). Within the Marsa catchment, the Lower Coralline Limestone is brought to the surface in several inliers, associated with faults. These are located around Attard and within the greater Siggiewi area. Occasionally silicified, the colour ranges from pure white to red and buff. This hard rock consists mainly of the remains of calcareous algae apart from molluscs, echinoid fragments and common foraminifera. Around Attard, where the limestone is silicified, is a particularly well developed algal horizon (ibid.). At or very close to the top of the Lower Coralline Limestone lies the Scutella bed, a layer with a great concentration of large flat seaurchins (“sand dollars”) belonging to this genus. A few feet thick, this layer is easily identified and defines the border to the overlying Globigerina Limestone (Pedley et al., 2002: 46). However, in some places, like around Attard, the Scutella bed cannot be recognised (House et al., 1961: 26). 2. Globigerina Limestone directly overlies the Lower Coralline Limestone and is the uppermost layer over large areas in western Gozo and central, southern and eastern Malta. As this limestone is softer, it weathers into gentle slopes and provides the building stone for most of the buildings on the islands. Varying in thickness between 20 6

m to over 200 m, the formation is comprised largely of fine-grained white to buff calcarenite full of calcareous tests of globigerinids, anomalinids and other small foraminiferans. Very thin horizons of phosphatic nodules have developed within the formation and two of these extend throughout the islands. This allows the subdivision of this formation into Lower, Middle and Upper Globigerina Limestones (ibid.: 27). The prevalent bedrock in the Marsa plain is Lower Globigerina Limestone (British Petroleum Exploration Company, 1957).

augite, feldspar and tourmaline. Due to its impermeability, this layer acts like a seal below the Upper Coralline Limestone and Greensand aquifer, and is thus the only horizon in Malta, where springs occur abundantly. This clay may have been used for much of Malta’s prehistoric pottery as suggested by the unbaked pottery vessel recovered from a Bronze Age context at Ghar Mirdum (MAR, 1964). 4. Greensand is a sandstone that consists of fossil debris with brown phosphatic grains and rounded grains of glauconite. Bright green when freshly exposed and unweathered, the glauconite turns orange-brown as it oxidises to limonite. It is the thinnest of all formations, generally not thicker than a meter, but can occasionally reach a thickness of around 15 m. As it does not significantly affect the form of the land surface, it is sometimes not recognised as a separate layer (e.g. Pedley et al., 2002: 36). The transition between the Greensand bed and the overlying Upper Coralline Limestone is difficult to trace, as the outcrops are limited in extent, but

3. The Blue Clay lies above the Globigerina Limestone, and may be up to 75 m thick, but at times thins out completely. There is a close lithological affinity as well as a striking resemblance of the fossil fauna with the underlying Globigerina Limestone, but morphologically, the horizon differs considerably from that of any other formation. It never forms cliffs or sheer drops but comparatively gentle slopes on which the clay is piled up in great heaps (Hyde, 1955: 43, 47). Within the clay, Murray (1870: 18) found rounded grains of quartz,

Figure 2.1.: Geology of Malta with Marsa catchment outlined. After House et al., 1961.

7

the lower transition towards the Blue Clay is generally clearly marked (Hyde, 1955: 49).

Sicily. During convergence processes, this formerly passive edge of the continent was transformed into a horizontal rift zone. Long lasting and extensive tensions along the Malta Escarpment resulted in the stretched structures with ridges (horsts) and valleys (graben) (ibid.: 9, 19) as a result of normal faulting. These geological faults cut through the rocks on Malta in east-northeasterly and north-easterly directions, the largest of them being the Great (Victoria Lines) Fault that crosses the island from Fomm ir-Rih on the west coast to the Madliena Tower on the east coast that effectively splits Malta into two regions. The landscape north of this fault is dominated by a series of rift and horst blocks that break up the northern part of Malta into a series of valleys and ridges. However, displacement on the faults is less conspicuous south of the Great Fault (House et al., 1961: 30).

5. The Upper Coralline Limestone is the youngest and topmost Tertiary formation and generally caps the highest hills. It is a hard, pale grey limestone that appears similar to the Lower Coralline Limestone. It also consists mainly of the remains of calcareous algae and forms sheer slopes (Pedley et al., 2002: 35). Within the Marsa catchment, the Upper Coralline Limestone is found together with the Blue Clay and Greensand formations in the Rabat/Dingli area, and is thus located at the fringe of the Marsa catchment (Figure 2.1.). 2.2. Tectonics – the structural setting of the Maltese Islands in the Strait of Sicily The Maltese Islands, together with the Hyblean Plateau on the Ragusa peninsula of southeast Sicily, belong in a plate-tectonic sense to Africa. The northern edge of the African plate is subdued in Sicily by the Maghrebides belt (Figure 2.2.). Together with southeast Sicily and the Hyblean Plateau, the Maltese Islands form a shelf unit on their own (Grasso & Reuther, 1988: 107), which is structurally, magnetically and gravimetrically distinct from the adjoining African foreland (Reuther et al., 2002: 7-8). The border of the African plate is located 100 km or so to the east of Malta (compare Pedley et al., 2002: 27, Fig. 22; Carminati & Doglioni, 2004: 2, Fig. 1), where the edge forms the Malta Escarpment. This escarpment runs in a NNW-SSE direction and shapes the east coast of

To the south of Malta, extending in a north-westerly direction, lie the Sicily Channel and the Pelagian shelf off the coast of eastern Tunisia. Both have been undergoing extension since at least Pliocene times, in the sense that Africa is moving south-westwards in relation to Sicily and the Maltese Islands. This rifting process leads to the deepening of the Pantelleria and Malta grabens (Carminati & Doglioni, 2004: 10). Its development began during the Quaternary in the south between Sicily and Tunisia and is still active today (Reuther et al., 2002: 13). The deepening of the grabens resulted in upwarped shoulders on the edges, which, to the west rose above sea level as the Pelagian Islands of Lampedusa and

Figure 2.2.: The Maltese Islands in relation to the present geodynamic framework. The active spreading in the Tyrrhenian sea and the opening of the Pantelleria Rift also created several active volcanoes. Source: Pedley et al., 2002: 27.

8

Lampione, and on the eastern shoulder as the Maltese Islands. Because of the upwarping shoulder, the Maltese Islands are tilting towards north-east (Reuther, 1984b: 14). Another result of this transformation system are minor fault lines that run parallel to this rift in a general NW-SE direction, of which the Maghlaq fault along the southern coast of Malta is the principal one (see Figure 2.3.). Here, this neotectonic Maghlaq Fault system plays its role as the easternmost master fault of the Pantelleria rift (Grasso et al., 1985: 16).

years ago, followed by passive graben infilling less than 1.5 million years ago. Both rift trends (the NW-SE Maghlaq Line direction and the ENE-WSW Great Fault direction) were generated in response to N-S stretching (Dart et al., 1993: 1153). House et al. (1961: 32) suggest a recent date for the Maghlaq Fault (Pleistocene) on account of Pleistocene molluscs that were caught up in a crushed conglomerate there. This is in agreement with the findings of Reuther et al. (2002: 13) who ascribe a Quaternary date to the roughly parallel running Pantelleria Rift. Lack of pronounced seismic activity in the Pantelleria Rift indicates that displacement originating from there occurs mainly by creep, while stresses in the north-east running Malta Escarpment may result in upthrusting (Reuther, 1987: 75). The faultings and mass movements are thus related to earthquakes with large magnitudes.

The normal faulting of the Maltese Islands is a result of tectonic movements, which appear to have deformed the Maltese Islands through the same stretching- and horizontal rifting processes that shaped the Gela Basin in the Strait of Sicily and the eastern edge of the Hyblean Plateau during the Plio-Pleistocene (Reuther et al., 2002: 8). Both the Malta Escarpment and the grabens within the Pantelleria Rift are part of a complex and interrelated tectonic system since the Upper Miocene to the Quaternary (Grasso et al., 1985: 17). Fault slip data from fully integrated analyses indicate that major fault activity occurred less than 5 million

The origin of many sheet fractures, particularly apparent in the Upper Coralline Limestone, may vary. Erosion of the underlying Blue Clay may be one reason that leads to fracturing and subsequent collapse (Pedley et al., 2002: 85), but compressional stress may also play a substantial

Figure 2.3.: Fault pattern of the Maltese Islands with the Marsa catchment outlined. After Reuther, 1984.

9

role (Twidale et al., 1996: 22). Fissures filled with red sediment may possibly also be ascribed to older than subrecent opening processes within the rock, although their age is often unclear. The width of these fissures, which occur in the Coralline and Globigerina Limestone between Marfa Ridge and Marsaskala, varies between a few centimetres to more than a meter, while they can reach a depth of several meters. However, as some fissures carrying land fauna fossils extend today below sea level (e.g. at Benghisa Point), these may indicate subsidence of a possibly more recent date (Vossmerbäumer, 1972: 50).

floor is around 100t per m² per 100 m depth, which can have a considerable effect on tectonic movements on the sea bed (Pirazzoli, 1998: 13). In the case of the Maltese Islands, it is tempting to presume a similar relative coastal stability as in the case of the southern Sicilian coast and Djerba, but there is no evidence that the odd sub-recent tectonic movement may not be attributed to the effects of varying sea levels. 2.2.2. Historical earthquakes Tectonic activity reaches up to the present. The study of historical seismicity of the Maltese Islands may provide important evidence about the spatial and temporal distribution of earthquakes. To establish with reasonable certainty how seismically active the area really was requires a long period of study. Before the 16th century there is no mention of seismic activity in the Maltese Islands in any of the surviving literary sources. This may be due to sources being lost or too little damage to be worth noting. The occurrence of an earthquake is not an unusual event in Italy and Sicily, and as such it was in antiquity not sufficient in itself to be noted by the ancient historians: the event had to be in some way ‘abnormal’ (Guidoboni et al., 1994: 26). Furthermore, ancient written sources generally only recorded earthquakes in specific circumstances or in connection with specific geographical areas. The surviving written evidence that mentions earthquakes in the Mediterranean suggests that as time progressed, disastrous events were even ‘demoted’. Thus, the failure of the sources to mention earthquakes is normal, and is also found in later periods (ibid.: 58).

2.2.1. Sub-recent tectonic movements The Maltese Islands should not be regarded as a single unit, but as an assemblage of individual blocks of various sizes that are characterised by different and divergent movements. As a result, regional changes of the landscape also show great local variations due to the nonuniform distribution and/or different emphasis of smallscale tectonics (ibid.: II). The dating of the visible results of tectonic movements is often difficult, unless they impact on a dated feature. For example, sub-recent faulting and mass movement may be indicated by the cart ruts entering the sea at St. Georges Bay in Birzebbugia1 (House et al.: 1961: 32) or cart ruts running across an inlet in St. Paul’s Bay (Hyde, 1955: 103). Likewise, with the help of interrupted cart ruts, Vossmerbäumer (1972: 89) dates a collapse that occurred south of Paradise Bay. Apart from the ongoing debate on the age of these ruts,2 their disappearance, submersion or destruction only provides a terminus post quem, which, despite its broad timeframe, places these events nevertheless within the more recent Holocene.

Both the Malta Escarpment and the Pantelleria Rift are tectonically active, but apart from the magnitude of the earthquakes originating from there, the distance of the Maltese Islands to the epicentre is important. The closer the epicentre, the stronger the earthquake is felt. In the absence of early historical data for the Maltese Islands, there are several documented events in Sicily, which, due to the close vicinity may also have affected the Maltese Islands.

To what extent the Maltese Islands could have been indirectly affected by post-glacial rebound that occurred on the European continent may possibly be inferred from data and research conducted in nearby Sicily and Djerba. Recent studies (Antonioli et al., 2003: 1-19 and Lambeck et al., 2004: 1567-1598) have demonstrated that while for the last 125 ka Northeast Sicily has been experiencing an uplift of around 1mm a¯¹, which, for the last 6000 years is calculated to be 1.4mm a¯¹, the rest of Sicily appears to have been stable throughout the later Quaternary. Evidence suggests that the uplift measured in northeastern Sicily is more likely controlled by plate processes, which hide the majority of the effects of glacio-hydro-isostatic adjustment. Generally, the main hydro-isostatic effect is the changing load on the sea floor as ice sheets grow or shrink. This leads to a subsiding sea floor during and after deglaciation on one hand, while on the other hand, larger land bodies may be uplifted by the deglaciation process as the heavy ice mass melts (Lambeck et al., 2004: 1581). The load carried by the sea

The earliest historical event is connected with the eruption of Mt Etna in the winter of 426-425 BC which is mentioned by Thucydides (3.116.1-3), Aristotle (Mete. 2.8.366 a) and Orosius (Hist. 2.18.6). According to Orosius, a very violent earthquake struck Sicily in 426 BC and eruptions of the Etna additionally destroyed fields and farms (Guidoboni et al., 1994: 122). Between 361 and June 363 AD, the historian Libanius (in Orosius, Hist. 18.292) mentions several destructive earthquakes that occurred in Libya as well as in Sicily: “… like a horse tossing its riders, she [Earth] has destroyed a great number of cities – many in Palestine, and all those in Libya. The greatest cities of Sicily lie in ruins …”. There is also archaeological evidence for damage around this time at the Roman villa at Piazza Armerina (ibid.: 260). The most frequently mentioned and debated earthquake of Late Antiquity is the one that occurred in the morning of 21 July 365 AD, which was described as a ‘universal’ earthquake, due to its magnitude and far-reaching effects.

1

At St. George’s Bay, the submersion of the cart ruts has also been assigned to rising sea levels (e.g. Trump, 2000: 96). 2 For example, Trump (2002: 264) dates them to the Bronze Age, as does Evans (1959: 190-191), while Bonanno (2005: 338-9) places the cart ruts in the Classical Age.

10

Its epicentre lay in Crete and resulted there in a single uplift event of more than 8m on the western coast. This resulted in an extensive seismic sea-wave that caused suffering to countless people in Sicily, Greece and Alexandria (ibid., 267-274). Considering the distance to Crete and the noted damage in Sicily, the resulting tsunami could also have reached the Maltese Islands from the east, possibly also damaging exposed areas. It is not uncommon that Greek earthquakes are felt in Malta; between 1997 and 2006 alone, this occurred on four occasions. Whether greater damage ensues or not is dependent on the depth at which these tremors occur (Galea in Grech, 2006a).3

and/or resulted in structural damage were noted. According to the records, 34 earthquakes were felt in Malta between 1537 and 1886. The most severe ones occurred in 1693 (see above), 1856, 1861 and 1886, all of which also resulted in structural damages to churches, palaces and houses, but no casualties. The 1856, 1861 and 1886 events all seem to have had their epicentre in the Sicily Channel, as the seismic activity occurred in swarms and the shocks were accompanied by a roaring sound from underground (n.a., 1886; Faurè 1913: 1071), which is typical for earthquakes in the Sicilian Straits (Barbano & Cosentino, 2003). Faurè (ibid.) lists another eight earthquakes between 1887 and 1911. Yet, the earthquake that occurred in 1908 with a magnitude above 7, and which was strong enough to destroy Messina leaving over 80,000 people dead, was not felt in Malta, possibly because its epicentre lay in Calabria (Azzaro et al., 1999: 7). Nonetheless, it created a tsunami, which reached a height of two meters by the time it reached the Maltese shoreline and entered the Grand Harbour (Faurè, 1913: 1065). Between 1900 and 2001, more than 24 earthquakes of varying magnitudes were noted within a 100km radius of the Maltese Islands, none of which resulted in casualties or other serious damage (Pedley et al., 2002: 33-4). Small magnitude earthquakes that are strong enough to alarm the population are a fairly regular occurrence in the Maltese Islands,5 but they usually do not exceed a magnitude of 4.5 on the logarithmic Richter scale. Many more go totally unnoticed by the population (ibid.). Although events like the 1693 earthquake are rare,6 they have a very destructive potential and are likely to have contributed to the shaping of the Maltese Islands.

An earthquake that occurred in eastern Sicily on 31 August 853 may also have had its effect on Malta, but the extent of damage it caused in Sicily is unknown (ibid.: 282-3). Although several other earthquakes are listed for Sicily in Guidiboni’s extensive catalogue, those that have taken place close to Reggio Calabria and in the Tyrrhenian Sea have been omitted here, due to the distance of the presumed site of the epicentre to the Maltese Islands. The Malta Escarpment appears to be the most likely source for the largest earthquakes (above magnitude 7.0), which hit the region in historical times. Among these are also the large earthquakes of 1169 and 1693, which caused vast amounts of damage and casualties in Catania (Azzaro & Barbano, 1999: 9). While the latter one also substantially damaged several buildings in the Maltese Islands, including the Mdina cathedral (Pedley et al, 2002: 33), there is no documentary evidence about any damage resulting from the 1169 earthquake, which may have been yet more disastrous as it had its epicentre even closer to Malta (Azzaro et al., 1999: 7, Fig. 1.1). The lack of records could be due to a low population density and a general dearth of any kind of documentation in Malta at that time. It is possible that both the 1169 and the 1693 earthquakes could have generated a tsunami, which then would have reached Malta on the side facing Sicily, but again, there is no literary evidence. However, Hyde (1955: 111) mentions in connection with an earthquake that happened on January 9 and January 11, 16924 that the sea at Xlendi in Gozo supposedly receded for a mile and then rushed back again, adding to the general destruction. The south-western location of Xlendi, however, would make an epicentre within the Malta Escarpment between Malta and Sicily in this context unlikely (see Figure 1.5.). Thus, the Xlendi tsunami should possibly be linked with another earthquake.

Several of the more recent earthquakes may also be noted for their attributed cliff collapses. Faurè (1913: 1072) lists a cliff collapse at Delimara near Marsaxlokk with the earthquake of 1856, and a cliff collapse at Qammieh near Marfa for an earthquake that happened in 1874. The collapse of an underground cavern at Qrendi, known as Il-Maqluba, is also possibly due to the shocks of an earthquake (n.a., 1890). However, whether the caving in of its ceiling should really be connected with an earthquake that supposedly hit Malta in 1343 is debatable. Here, the dating of the event by Castagna, and after him Faurè, may be wrongly attributed (Pace Abela, no date: 1519), while a terrible storm accompanied by tremors in the night of November 24, 1343 is mentioned by the Italian poet Petrarch in a letter to Cardinal Colonna (Pace Abela, no date: 1518). This known event happened in Naples and appears to have been later linked with the collapse of the cave roof. This throws some doubt on the dating of the Maqluba collapse, which Hyde (1955: 20) describes as a “simple caving in of the surface above subterranean caverns”.

It appears that the recording of earthquakes started only after the arrival of the Knights of St. John, when, in the absence of advanced instrumentation for the detection of seismic activity, earthquakes that alarmed the population

The above written records of seismic events, which start for the Maltese Islands only in the 16th century, are unlikely to be complete. It was mainly the seismic

3 Two Greek earthquakes were felt in the Maltese Islands in 1997, one in 2002 and most recently, one in January, 2006. 4 Hyde possibly here refers to the earthquakes that happened in 1693 on the very same dates.

5 In recent years these occurred for example in 1990, 1992, 2001, 2003, 2006. 6 The return period for earthquakes with a magnitude > 7.0 in the Malta Escarpment is estimated to be 500 years (Moroni et al., 1999: 119)

11

activity of the Malta Escarpment and the Pantelleria Rift that has roughly shaped the topography of the Maltese Islands. Earthquakes that had their epicentres in other parts of the Mediterranean may also have affected the islands, as the recent earthquakes around Crete suggest (see above). Tectonic activity has been an ongoing process since at least the Quaternary and evidence suggests that earthquakes have been affecting the Maltese Islands ever since, despite the absence of any written records before the 16th century.

1961: 63, 66). The summers are described as hot and dry, and the winters as mild and wet (e.g. Nehring, 1966: 17; Chetcuti et al., 1992), with a mean monthly temperature of 12-26ºC (Schembri, 1997: 1). Temperatures at present do not fall long enough to such low levels that they would affect vegetation growth, but they can occasionally reach -5ºC at night during the coldest months. In summer, maximum temperatures at grass level have been reported to reach up to 49ºC, and at soil surface level they frequently are above 40ºC, which may be lethal to vegetation, especially seedlings (Haslam et al., 1977: xviii). The islands are apart from very sunny also very windy (Schembri, 1997: 1), with, over the year as a whole, only one day in ten that has no measurable air movement (Mitchell & Dewdney, 1961: 53). Relative humidity is mostly between 65-80 % and usually much higher than in neighbouring Mediterranean coastlands.

2.2.3. Tectonics in the Marsa area The geological faults within the Marsa catchment (see Figure 2.3.) appear to be mainly related to the drainage pattern and run roughly parallel to the Great Fault (see below and compare Figures 2.9 and 2.11). A geological fault at the Marsa bay, however, stands out. Aligned in the same direction of the Maghlaq Fault, it may perhaps be connected with the submersion of river valleys and the general tilting towards the north-east. This rather minor fault does not feature in many maps that show the faulting, but its alignment with the Maghlaq Fault possibly suggests a comparatively recent date (Reuther, 1984: 74).

Often, combinations of wind, temperature and humidity have a negative effect on plant growth. Particularly the hot, humid Xlokk, the cool north-westerly Majjistral and the dry north-easterly Grigal affect transpiration rates and wilting points to a considerable degree (ibid.). Like the rest of the Mediterranean, the islands have experienced marked climatic fluctuations, with periods of persistent drought and wetter, cooler conditions (Hughes, 1999: 62; Grove & Rackham, 2001: 133).

The fairly straight opening of the Marsa embayment facing the open sea due north-east makes the area vulnerable to tidal waves that are the result of earthquakes, particularly when the epicentre is located north-east of Malta within the Malta Escarpment. The only recorded tsunami that entered the Grand Harbour was connected with the Messina earthquake of 1908, but also here, an absence of written records is by no means evidence of absence.

2.3.1. Climate and weather effects in the Marsa Plain in Modern Times Due to the size of the plain and its surrounding mainly gentle slopes (see below), the area is exposed to the elements. Furthermore, the rainwater catchment of the Marsa plain receives runoff water of an area covering 52.1 km², which is the largest rainwater catchment of the Maltese Islands.7 Water runoff is roughly 20% for rural areas8 and assuming that the Marsa catchment was primarily a rural area in the past, in a heavy (but not unusual) rainfall of 80mm within several hours it is estimated that some 708,560 cubic metres of water are washed into the Marsa basin.9 This kind of rainfall is no rarity especially in October and November, and is a hallmark of the Mediterranean climate (Grove & Rackham, 2001: 25-35; Perry, 1997: 31-33). Once any kind of river basin is full and the soils are saturated with water, additional rainfall results in flooding. In the past, a large part of this runoff water collected in the valleys or widien, namely Wied is-Sewda and Wied il-Kbir (see below, Figure 2.10.). Other seasonal tributaries channelled water into these widien all along the way down to the sea. Blouet (1964a: 198) reports that in the

2.3. Climate The climate of an area is a statistical notion defined over 30 years by the average values of four meteorological parameters: precipitation, cloudiness, temperature and wind. When these recorded values change significantly, we may speak of climatic change (Petit-Maire, 2003: 17). The present climate of the Maltese Islands may differ considerably from the prevalent climate earlier in the Holocene and particularly in the Late Pleistocene. It serves, however, as a gauge against which the different climatic regimes of the past may be measured. Climatic data started being measured since 1841, but many early measurements are lost. Reliable data started being measured at an increasing number of stations around Malta as from 1884 (Mitchell & Dewdney, 1961: 48, 58). The present climate of the Maltese Islands can be described as typically bi-seasonally Mediterranean, with an average rainfall of around 530 mm, 85% of which fall between October and March, but the precipitation is highly variable from year to year (Chetcuti et al., 1992). Storm rainfalls are frequent, and may be quite localised. In some years, parts of the islands may receive as much rain in 48 hours as they have, on other occasions, received in a whole rainy season (Mitchell & Dewdney,

7

In comparison, the Burmarrad catchment is 39.4 km² and that of Msida, 12.8 km². 8 I am grateful to Dr John Mangion of the Water Services Corporation, Luqa, Malta for providing this information. 9 In September 2003, more than 250mm of rainwater fell within less than 24hrs. This resulted in catastrophic floods, especially at Qormi and Marsa, where over 2 150 000 cubic meters of water found their way into the Marsa plain. Strong rainfalls have always occurred periodically: Mitchell & Dewdney (1961: 64ff) describe in detail storm rainfalls in October 1913 and November 1915, where precipitation exceeded even 400mm in places!

12

17th century large quantities of debris washed down with the torrents produced by Malta’s heavy winter storms would block the lower courses of the widien and create stagnant pools, unless the drainage channels were regularly maintained. This situation has not changed much, although today an artificial canal channels the rainwater from Wied is-Sewda and Wied il-Kbir into the sea.

2.3.2. Factors inducing climatic change Various occurrences may influence the climate and weather. On a short scale of one to a few years, volcanic activity may, through throwing masses of ash into the atmosphere, cool certain areas if the stratosphere is reached. Here, the vicinity of Mt Etna to the Maltese Islands may be significant. Although the Mt Etna eruptions in 2002 may be described as spectacular (Pedley et al., 2002: 13), the ash plume that reached the Maltese Islands only resulted in a very thin ash layer, with no apparent consequences on climate and weather. On the other hand, a 30-45cm thick layer of volcanic tuff that was found to cover a Pleistocene deposit in a hollow at Mriehel (Zammit Maempel, 1989: 44) might point to a significant volcanic event at an undated point in the Pleistocene past, and this could have had considerable consequences on the local weather and vegetation.

Wind is also a major factor affecting the Marsa Plain as the hills are not high enough to shelter the area significantly. The main direction of wind for the Maltese Islands is from the north-west, followed by winds from a westerly direction. These two directions also have the highest occurrences of wind speeds exceeding 11 knots (Department of Civil Aviation, 1988). For the Marsa plain this means that if these north-westerly and westerly winds are strong, the valley slopes may act like funnels and increase the speed of the wind. That the plain is windswept may perhaps also be deduced from maps of Malta from the 16th and 17th century: many depict the Marsa Hortus – (a tree orchard) with a high wall around it.10 Apart from providing security against possible intruders, the high wall would have also protected against the effects of strong wind.

Climatic variations on a middle scale are mainly attributed to solar activity (recorded by sunspots and auroras) and induce moderate climatic changes lasting one to several centuries (Petit-Maire, 2003: 17). Examples here are the Medieval Warm Period, which lasted from the 10th to the 13th century and which allowed, among other, the Vikings to colonise Greenland and vine to grow in England (Bojanowski, 2003), and the Little Ice Age, a period of extreme weather that followed the Medieval Warm Period and which lasted in the Mediterranean until the 1710s (Grove & Rackham, 2001: 137). However, palaeoclimatic records from ice cores, deep sea and lake cores demonstrate that climatic conditions were highly variable throughout the Holocene (e.g. Robinson, 1994: 235; Combourieu Nebout et al., 2002; Sangiorgi et al., 2003b) and more so during the last glacial (e.g. Combourieu Nebout et al., 2002: 863-4 and Figure 2.4.).

To the north-east, the area opens to the sea. Here, the north-easterly Grigal wind blows straight into the mouth of the Grand Harbour and, depending on the former shore topography, may have raised waves high enough during storm surges to sweep the shoreline and possibly caused damage to any agricultural soil and produce that lay within reach of spray and surge. Faurè (1913) provides a long list of weather phenomena dating back to the 15th century. The list is likely to be biased, as, similar to the list of earthquakes mentioned above, the considerable damage the weather caused made it noteworthy. By far the most mentioned damage is connected with the Grigal storms, as it regularly damaged ships in the Grand Harbour. Although today a breakwater protects the inner part of the Grand Harbour, it does not prevent ships from breaking their mooring when wind gusts reach Force 11, as happened on February 1, 2006 (Testa, 2006). The wind also drives material onto the foreshores and may thus create humped beaches, which act as barrages against the storm water. During violent storms the barrage may be breached or overridden and the land behind flooded. This reportedly happened several times during the 17th century (Blouet, 1964a: 199). As the Grigal wind blows on 15% of all days on an average year, its effect and impact on the Marsa plain may be quite considerable (Mitchell & Dewdney, 1961: 53).

In contrast, at the geological millennial scale, global climate periodically undergoes major shifts between a generally warm, ice-free state and an ice age state with continental ice sheets. Through geological time, there have been at least seven major ice ages. These are determined by a combination of solar luminosity, continental location, tectonics and atmospheric CO2 concentration (Barry & Chorley, 1998: 330) while the alternation is a result of pseudo-periodic changes in the Earth’s orbit (Petit-Maire, ibid.). The last glacial period is dated to around 125 ka ago, finding its climax in the Last Glacial Maximum (LGM) around 20 000 years ago. ‘Modern’ climatic conditions became established only around 10 000 years ago, marking the start of the Holocene (Barry & Chorley, 1998: 331).

The gentle slopes that surround the Marsa Plain offer no significant protection from the sun either. Thus, as mentioned above, the adverse effect on plant growth resulting from a combination of elevated temperatures, high wind speeds and low humidity conditions, may be particularly pronounced in the Marsa Plain.

2.3.3. Past climates of the Maltese Islands As dated material from the sediments of Marsa Core 1 spans between around 45,000 years bp11 to around mid 11 The letters cal. BP following a date indicates a calibrated radiocarbon date expressed in calendar years ‘before present’ (i.e. 1950), while bp indicates uncalibrated radiocarbon years.

10

For example, Johannes Quintinus’s map of Malta, published in his book Insulae Melitae Descriptio, Lyons, 1536. See Figure 2.8.

13

Figure 2.4.: Comparison between GISP 2 record and palaeoenvironmental reconstructions from Ocean Drilling Program (ODP) Site 976. Interstadials are indicated along Site 976 through graphical correlation with Greenland ice core record. YD = Younger Dryas, H1- H5 are Heinrich events. Source: Combourieu Nebout et al., 2002

15th century AD (estimated, see below), the possible climate of these periods only will be discussed. In broad terms, this time span encompasses the middle Late Pleistocene to the Holocene, or from around the middle of Oxygen Isotope Stage 3 to Oxygen Isotope Stage 1, and includes several warming and cooling events. For the Maltese Islands, broad palaeoenvironmetal reconstructions have been made on the basis of the faunal record (e.g. Trechmann, 1938; Storch, 1974; Fischer & Stephan, 1974; Hunt & Schembri, 1999; Borg, 1999), but they lack in absolute dates. As such, the findings are quite low resolution, but still allow a broad picture to be drawn. Storch’s faunal analyses that reach from the Early Pleistocene to the Bronze Age allowed him to conclude that all three faunal stages analysed indicate a Mediterranean climatic regime (1974: 429).

sediments of Lampedusa that date back to around 41,000 bp (Giraudi, 2004: 537-545). These findings may also be supplemented with the continuous high resolution proxy records from Site 976 in the Alboran Sea in the western Mediterranean and the palaeoclimatic records of the GISP2 ice-core record from Greenland (in Combourieu Nebout et al., 2002: 863-866). Correlation of the various records shows that despite the wide geographical range, the results are comparable (see Figure 2.4. and below). This emphasises that the Northern Hemisphere oceanatmosphere system of the last glacial period extended its influence at least as far as the central Mediterranean region, as the Monticchio and Lampedusa records demonstrate. For the transitional period to the Holocene and for the Holocene itself, available data from the Sicily, Tunisia and core SD91-17 from the southern Adriatic Sea (eastern Mediterranean) are added to deduce a possible climatic picture for the Maltese Islands for this time (see Figure 2.5.).

There are, as yet, only a few proxy records that reach back to the Pleistocene from the central Mediterranean. Most important here is the lacustrine sequence from Lago Grande di Monticchio in the Basilicata region of southern Italy,12 which extends 102,000 years back (Allen et al., 1999: 740-743), and the Upper Pleistocene to Holocene

2.3.4. Climate from ca 45,000 bp to the Late Glacial at around 15,000 cal. BP This 30,000 year long period within last glacial is marked by numerous rapid climate fluctuations as indicated by the oxygen-isotope records from Greenland ice cores,

12

The lake lies around 600m asl in a nature reserve of the Appenine Region and is surrounded by a dense forest.

14

another significant wet event occurred (Chondrogianni et al., 2004: 37). During the less cold or dry phases on Lampedusa, soils developed concurrently also in Tunisia around 19,800 +/- 500 bp (Coudé-Gossen et al., 1987) and on Jebel Gharbi in Libya in 21,610-20,090 cal. BP (Barich & Giraudi, forthcoming), indicating a moderately dry climate (Giraudi, 2004: 541). At Lake Albano, a lag deposit at around 15,200 cal. BP indicates an extreme lake level lowering and increased erosion, which may be attributed to a brief cooling episode known as the Oldest Dryas and documented in several Mediterranean marine records (Chondrogianni et al., 2004: 37). Sea surface temperatures in the Sicily Channel (derived from alkenone analyses) only reach between 11ºC and 15ºC in the late Pleistocene (Sangiorgi et al., 2003a).

Figure 2.6.: Location of the various sites with palaeoclimatic data discussed in the text. 1 = ODP Site 976 (Combourieu Nebout et al., 2002), 2 = Medjerda Plain, Tunisia (Zielhofer et al., 2004), 3 = Lampedusa, Italy (Giraudi, 2004), 4 = CS2001 (Sangiorgi et al., 2003a), 5 = Lago di Pergusa, Sicily (Sadori & Narcisi, 2001), 6 = Lago Grande di Monticchio, Italy (Huntley et al., 1999), 7 = Lago di Mezzano, Italy (Sadori et al., 2004), 8 = Core AD91-17 (Sangiorgi et al., 2003b.

For the Maltese Islands, the environment during the Late Pleistocene is described as open, exposed and lacking in vegetation with deposition of loess, and severe flash floods that would have deposited fluvial and colluvial sediments (Hunt, 1997: 107; Schembri & Hunt, forthcoming), while the extinction of the Late Pleistocene endemic vole Pitymys melitensis before or with the onset of the Holocene is seen as indicative of the disappearance of deep soils and/or too severe seasonal aridification (Malec & Storch, 1970: 78). An analysis of the habitats of the Pleistocene avifauna discovered in the Maltese Islands reveals that only four of the 39 species listed and identified prefer a wooded habitat, while an overwhelming number preferred open shallow waters (brackish and/or fresh) and/or steppe (Borg, 1999: 77-89). Although it is not known to which period or periods within the Pleistocene all these bird bones belonged, the small number of birds associated with forests or woodland is, however, suggestive. Furthermore, of these four bird species, three are migratory birds while only one may possibly have been sedentary (ibid.).

variabilities in sea surface temperatures as evidenced in North Atlantic marine sediment cores, and markedly low, but fluctuating eustatic sea level variations. The cores from Lago Grande di Monticchio indicate that the accompanying vegetational changes frequently occurred in less than 200 years, and that decreases in taxa happened typically more quickly than increases (Allen et al., 1999: 740-741). Data from Monticchio also indicate that moisture was reduced during the cold or warm steppe periods within the glacial and extreme temperatures were up to 20ºC below modern equivalents during cold minima. On the other hand, during the rapid climate fluctuations, increases in the mean temperature of the coldest month in particular exhibits changes of more than 10ºC within a mean interval of 153 years.13 Towards the LGM at around 20,000 BP, the Lago Grande di Monticchio pollen record shows predominantly open ground, with only little pollen of pine and juniper, and the absence of birch, oak, alder, ash, beech and any other woody taxa (ibid.: 741).

2.3.5. Climate from the Late Glacial at around 14,500 BP to the onset of the Holocene (around 11,400 BP)

On Lampedusa, the period extending between 40,840 +/720 bp to more recent than 35,960 +/- 1300 bp may have been a period marked by strong winds, as indicated by the deposition of Saharan dust (Giraudi, 2004: 540). Soils developed with calcareous concretions and carbonate crusts under wetter conditions between 30,190 +/- 280 bp and 28,920 +/- 560 bp, a trend that could also be observed in Tunisia and Libya at about this time. A wet event around 28,000 bp has also been identified at Lake Albano in central Italy (Chondrogianni et al, 2004: 37). Towards the LGM, a marked phase of erosion took place on Lampedusa, caused by a considerable lowering of the sea level. Here, a reduction of the vegetation cover may have been the cause for an increase in the production of slope debris, and thus the climate would have been drier or colder than around 23,590 +/- 110 bp, when the deposition of Saharan loess prevailed (Giraudi, 2004: 541). Around the same time, the lake levels of Lake Albano in central Italy rose around 30m, indicating that

From around 14,500 BP onwards, the climate shows signs of amelioration as temperatures rise and humidity increases. Sea surface temperatures in the Sicily Channel increase by around 3ºC and reach around 18ºC (Sangiorgi et al., 2003a). This wetter and warmer period is known as the Bølling/Allerød and it is not only recorded in central Europe, but is also reflected in lake sediments in the central Mediterranean region (e.g. Allen et al., 1999: 741; Huntley et al., 1999: 956; Ramrath et al., 2000: 89) and in the stratigraphy of the Medjerda floodplain in Tunisia (Zielhofer et al., 2004: 856-7). Here, an analysis of the macroremains indicates that bioclimatic conditions were already similar to the ones of today (ibid.). Similarly, at Lago Grande di Monticchio, the palaeoclimate deduced from pollen records also points to conditions comparable to the Holocene (Allen et al., 1999: 742). This warming event was interrupted by a brief cold period between 13,250 and 13,070 BP, to which the woody taxa at Lago di Monticchio reacted by a sharp decline in Quercus pollen, complemented by a sharp rise

13 For comparison, the mean annual temperature in Malta increased over the last 77 years by 0.5ºC only (Grech, 2006b).

15

in Betula. All in all, the Bølling/Allerød lasted until 12,650 BP at Lago di Mezzano in central Italy (Ramrath et al., 2000: 89) and until around 12,400 BP in Tunisia (Zielhofer, 2004: 856) and was then followed by a rapid deterioration of the climate, with marked cooling and clear aridification, known as the Younger Dryas. Similar to the previous Bølling/Allerød period, the Younger Dryas is also seen as a large-scale climate event of Northern Hemispheric extent that lasted between 1000 and 1500 years according to the Greenland Ice Core and some lacustrine deposits (e.g. Ramrath et al., 2000: 89), but only 600 years in Tunisia (Zielhofer, 2004: 857).

in Sicily (Sadori & Narcisi, 2001: 11), at Lago di Mezzano in Central Italy (Ramrath et al., 2000: 89), and also in the southern Adriatic Sea, as reflected in the pollen diagrams from core AD91-17 (Sangiorgi et al., 2003b: 727) before reafforestation starts. The start of postglacial reafforestation is set at at Lago di Pergusa at around 12,950-11,750 cal. BP, when the percentage of arboreal pollen starts to increase (Sadori & Narcisi, 2001: 663). Slightly later, by 10,000 BP open ground taxa at Lago Grande di Monticchio were slowly overshadowed by predominantly Quercus (Allen et al., 1999: 741). Between 10,450-9700 cal. BP a significant change took place in the Central Mediterranean as numerous palaeoclimatic findings indicate a thermal and hygric climate optimum. On a larger scale, the findings correlate with an overall humid period in the western and eastern Sahara, Sahelian and central Arabian region (Zielhofer, 2004: 858 and references therein) as well as with the records from core AD91-17 from the southern Adriatic Sea (Sangiorgi et al., 2003b and see Figure 2.6.). This thermal and hygric climate optimum leads to a dense forest cover in many areas of the central Mediterranean, characterised by the dominance of deciduous and evergreen oaks (e.g. Lago di Pergusa, Lago Grande di Monticchio, Lago di Mezzano, Ionian Sea cores RC 9191 and KS 50) and to the deposition of the sapropel layer S1 in the Adriatic Sea as a result of suboxic and anoxic conditions caused by increased algal blooms (Sangiorgi et al., 2003b). By 8,180-7,930 cal. BP, at Lago di Pergusa in Sicily, Olea, Ulmus and Pistacia begin to expand following a drop in Fagus, Corylus and

Despite the possible variations of the durations of these events, it is very likely that they would also have affected the Maltese Islands by a rising sea level and change or increase in vegetation. Through the cooling and aridification event of the Younger Dryas these could have, however, again reversed the vegetational cover to possibly close to Late Glacial levels with a predominantly steppe environment. In the absence of any carbon dates that can be closely related to the environment this remains, however, speculation. 2.3.6. Early Holocene During the Postglacial, the climate in Sicily appears to be marked by relatively arid conditions (Sadori & Narcisi, 2001: 11), while in Tunisia, a humid climate set in (Zielhofer, 2004: 858). Thus, at the onset of the Holocene at the end of the Younger Dryas at around 13,850-12,950 cal. BP, a predominantly steppe environment is indicated

Roman Optimum mid-Holocene climatic collapse P2 cooling event Climatic Optimum

Cooling event

Younger Dryas

Figure 2.6: Down-core records of different sea surface temperature (SST) proxies in core AD91-17 from the Southern Adriatic Sea. From left to right: oxygen isotopes measured on G. bulloides; relative abundance of warm-water dinocyst species calculated on the oxygen resistand dinocyst sum; alkenone-based SSt; number of Syracosphaera pulchra per mm²; relative abundance of warm water foraminifer; relative abundance of decidous Quercus, semi-desert pollen and Pistacia. Error bars for oxygen isotopes and for the alkenone-based temperatures are based on the analytical standard deviation. Other error bars not available. The darker grey horzontal band indicates sapropel S1, the lighter band within, its interruption. The two upper lighter grey bands indicate different Adriatic cooling events. Source: Sangiorgi et al., 2003b: 727.

16

Ericaceae, but whether the drop is due to climatic factors or to anthropogenic action cannot be ascertained (Sadori & Narcisi, 2001: 668).

occurred for about two millennia, during which widespread afforestation took place. The aridification trend that started in 8,350 cal. BP in Sicily was overlapped by human impact on the vegetation, but the human impact is not said to have produced devastating effects on an already open landscape (Sadori & Narcisi, 2001: 670). Alternating oscillations in hot-arid and hot-humid climatic conditions, which prevailed throughout the Late Pleistocene/early Holocene left their mark also in the deposits outcropping along the Hyblean coastal zone of south-east Sicily (Palomba & Tedesco, 2001: 50). A hot climate is also indicated by sea surface temperatures in the Sicily Channel, which reach up to 20ºC between 10,000 and 6,000 BP (Sangiorgi et al., 2003a).

An early to mid-Holocene abrupt cold event documented for the central Mediterranean (Tunisia and Lago di Mezzano) as well as in core AD91-17 from the southern Adriatic Sea at around 8,400 to 8,000 BP possibly also involved an aridification of the climate. The event is marked by geomagnetic variations and in geomorphic profiles (Zielhofer, 2004: 858; Ramrath et al., 2000: 901), and is also reflected in pollen profiles at Lago di Pergusa in Sicily (see Sadori & Narcisi, 2001: 668). Furthermore, the pollen profile from AD91-17 shows a marked decrease in deciduous oaks (see Figure 2.6.).

In the Maltese Islands, the nature of the early Holocene environment is very poorly known. Due to the abovementioned afforestation that took place in much of the central Mediterranean as a result of increased humidity and warmer temperatures, an evergreen forest dominated by oak and pine is often assumed for the Maltese environment (e.g. Schembri & Lanfranco, 1993; Grech, 2001; Schembri & Hunt, forthcoming). This potential past existence of evergreen forest, however, and its subsequent disappearance needs scientific assessment, as would the reason and date for its postulated disappearance. Like in Sicily, any forest in Malta may have opened up considerably before the arrival of the first settlers, after the climate had become more arid, but a forest may also have been cleared by the first settlers (e.g. Grech, 2001; Schembri, 1993; Giusti et al., 1995; Trump, 2002). Environmental data from the excavations at the Xaghra Circle in Gozo (Figure 1.5.) indicate that in the earliest phase represented (Zebbug Phase, from ca. 4350 cal. BC) the site was treeless, but covered with relatively dense vegetation of a steppe or garrigue aspect (Schembri & Hunt, forthcoming). The continuous thinning of the vegetation and exposure of limestone during the subsequent Ggantija and Tarxien Phases (between around 3,360 cal. BC to around 2,500 cal. BC) may be the result of human impact (ibid.), possibly coupled with the aridification of the climate.

Another event, possibly more severe than the previous one, is indicated by the deposition of coarser sediments in the Medjerda overbank sediments between 6,600 and 6,000 BP, pointing to drier conditions in Northern Tunisia (Zielhofer, 2004: 858). This may also be indicated at Lago di Mezzano, where between 6800 to 6200 varve years BP, the arboreal pollen percentage decreases from 95% to 80%, combined with a drastic decrease in the total pollen concentration (Ramrath et al., 2000: 91). A possible cause for the decrease in the pollen concentration could be a drop in the mean precipitation that in turn caused decreased runoff as well as restricted plant growth in the catchment area (ibid.). The aridification event, also known as P2, may also be accompanied by marked cooling, as indicated by a decrease in winter sea-surface temperatures by 2ºC in the southern Adriatic Sea at around 6,000 BP (Sangiorgi et al., 2003b: 723 and Figure 2.6.). At Lago di Pergusa in Sicily, a moisture decrease is noted from around 8,350 cal. BP onwards, to which the vegetation composition responds by 8,150 cal. BP and the biomass by 7,540 cal. BP. The climatic trend towards aridification continues progressively to about 6,000 BP and apart from a brief amelioration at around 5,350 cal. BP continues getting more and more arid (Sadori & Narcisi, 2001: 669). Conditions are favourable for soil formation in Tunisia between 6,500 to 5,000 cal. BP, indicating a relatively stable and warm climate with sufficient humidity (Zielhofer, 2004: 859). At Dar Fatma in Tunisia, after ca. 5,250 cal. BP, changes in the arboreal taxa also clearly indicate a trend towards more arid conditions (Ben Tiba & Reille, 1982). Severe aridification of the climate at 4,700 BP is also indicated in the Medjerda floodplain in Tunisia (Zielhofer, 2004: 859), at Lago di Mezzano (Sadori & Narcisi, 2001: 669) and in the core records from the southern Adriatic Sea (see Figure 2.6.). This corresponds with other results from the Mediterranean, but also from the western and central Sahara, where drastic aridification that led to the present ecological conditions has been recorded between 5,500 and 4,500 cal. BP (Zielhofer, 2004: 859).

2.3.7. Late Holocene (around 4500 BP to present), central Mediterranean and Malta The ongoing aridification trend that started at around 8,350 cal. BP in Sicily continued and had its peak in 6,650 cal. BP resulted in a markedly more open wood structure at Lago di Pergusa from around 5,000 cal. BP. Even drier conditions started at around 3,500 cal. BP and lasted until about 2,900 cal. BP, when the climate became moderately dry. For the last 2000 years, the sediments of Lago di Pergusa do not indicate any pronounced climatic change (Sadori & Narcisi, 2001: 670). On Lampedusa, greater aridity is reflected in the sediments at around 4,560-4,410 cal. BP. Around 3,000-2,750 cal. BP there was a short wet climatic phase that favoured the growth of a small stalactite in a cave in Forbice Valley on Lampedusa (Giraudi, 2004:542), which coincides with the increase in precipitation in Sicily.

As records indicate, the Postglacial climate in the central Mediterranean appears to be characterised by relatively arid conditions, while truly wet conditions only had 17

At Lago di Mezzano in central Italy, the picture is slightly different as here there is a marked deterioration of the climate between 5,000 and 4,000 cal. BP, which is also reflected in several other sediment archives and regions all over Europe. Increases in precipitation only start again at around 3,300 cal. BP (Ramrath et al., 2000: 91).

Schembri & Hunt, forthcoming) conformably indicate predominantly open environments in various periods, the data are too low in resolution to allow refined climatic deductions. However, the close conformities among the central Mediterranean data listed above make it likely that the changes that occurred in nearby Tunisia, Lampedusa and Sicily in particular, would also have affected the Maltese Islands. As from the 15th century, there is some scanty literary evidence with regards to the weather in Faurè (1913), who lists some weather phenomena up to 1908, covering more than 500 years. From his account, it appears that heavy storms and rainstorms were a very common occurrence that often resulted in great damage in the Grand Harbour, but also in the countryside. Hurricanes were infrequent, but excessive cold weather is mentioned several times, while severe droughts, with sometimes no rain at all are said to have occurred nearly every century. There is a considerable increase in Faurè’s data as from 1800, which is probably more likely to be due to increased recording rather than an increase in weather phenomena. Because of this possible bias, it is difficult to interpret the earlier records as results of the Little Ice Age, as it appears that erratic weather occurred throughout these 500 years.

In the southern Adriatic Sea, there is a marked decrease in sea surface temperatures of more than 2º C between 4,000 BP and 3,000 cal. BP, after which they rise again. (Sangiorgi et al., 2003b: 727 and Figure 2.6.) Data from the historical period from the Hyblean coastal zone in south-eastern Sicily indicate cold-humid conditions between 520-350 BC (Little Archaic Ice Age), between 500-750 AD (Early Middle Ages) and between 1500-1850 AD (Little Ice Age), and a hot-arid period in the Middle Ages between 1000-1300 AD (Palomba & Tedesco, 2001: 51). Cooling is also indicated by the sea surface temperatures from the Sicily Channel, which decreased from 20ºC between 10,000 and 6,000 BP gradually to reach 16ºC by 900 BP (Sangiorgi et al., 2003a). At the Medjerda floodplain in Tunisia, the abrupt midHolocene aridification that started at 4,700 BP lasted until around 3,600 BP. At around 3,000 BP drier conditions set in again that are possibly related to an abrupt desiccation phase that occurred widely in the western Mediterranean (Faust et al., 2004:1770). Humid conditions probably first set in at around 2,500 BP, which increased land use for agricultural purposes as more land became suitable for cultivation by the Punic people. Land use and exploitation was enhanced after 146 BC by the Romans. From the Late Roman period, climatic as well as human activities led to erosion and major floods occurred in the Medjerda floodplain, as more arid conditions set in around 1,700 BP that lasted around 300 years. These difficult climatic conditions for agriculture coincide with the social decline in the Mediterranean that started with the invasion of the Vandals (ibid., 1771). The same trends may be observed from the records of core AD91-17 from the southern Adriatic Sea (see Figure 2.6.). By 1,000 BP, a period that coincides with the Oort sunspot minimum (cf. Fig. 4 in Faust et al., 2004: 1766), the climate is marked by a cool and dry period with increased erosion. The Medieval Warm Period (between 900 and 700 BP) corresponds to a period of stable sedimentation in North Tunisia, marked by globally rising temperatures and more humidity. More arid conditions mark the period between 700 to 500 BP, coinciding with the onset of the Little Ice Age. By 400 BP, disastrous flood events set in, which, on a global scale, correspond to the Maunder Minimum. The period between 1600 and 1830, is the coldest of the Little Ice Age (ibid.: 1772).

2.4. Geomorphology 2.4.1. General Despite the relatively small size of the Maltese Islands, there is great diversity in relief and landform, which is mainly the result of the effect of climate and tectonics on the geology, while human agency has changed the coastline in several parts of Malta.14 Tectonics resulted in the broad formation of the relief and landforms, while the actions of erosion by wind, rain, sun and sea refine and continuously reshape the landscape. In broad terms, the following major elements make up the physical landscape of Malta (after Dewdney, 1961a: 34 and Nehring, 1966: 9):15 1. The Coralline Limestone plateaux in the west, which form the highest parts of Malta and are bounded by well-marked escarpments (see Figure 2.1.). 2. Blue Clay slopes that separate the plateau uplands from the surrounding areas. They also occur in valleys at the plateau edges. 3. Undercliff areas known as rdum, which occur where the coralline plateau meet the sea. 4. The horst-and-graben landscape north of the Victoria Lines 5. Flat-floored basins. In most cases they are the result of faulting (e.g. Pwales Valley), 14

For example in the Grand Harbour (see below), but most recently also at St. Julians during the construction of the new Hilton Hotel where a yacht marina had been artificially constructed that effectively changed the coastline. 15 As the geomorphology of Gozo and Comino has little bearing on the present study, it is omitted here. For a detailed description of the geomorphology of Gozo and Comino, the reader is referred to e.g. Bowen-Jones et al.,. 1961; Nehring, 1966 and Pedley et al., 2002.

For the Maltese Islands, there are little environmental data for the Late Holocene. Pollen analysis from a Bronze Age cistern at Tal-Mejtin in Luqa indicates predominantly open ground (Trump, 1966), as does the pollen assemblage from a Punic pit fill at Tas-Silg (Hunt, 2000). But while pollen and molluscan analyses (see 18

6.

downwarping (e.g. Bingemma Basin), but they may also be due to erosion and subsequent alluvial deposition (e.g. Wied il-Ghasel). Globigerina hills and plains. These are large areas of gently sloping land that take the form of a series of low ridges and shallow valleys in the south and south-east.

plateau is deeply incised by Wied il-Busbies, which divides the plateau into the Bingemma Plateau and the Rabat-Dingli Plateau, and Wied Liemu. These two widien form, with other minor tributaries, Wied il-Ghasel, which eventually drains north into Salina Bay. Similarly, further south along this boundary, the tributaries that lead eventually into Wied is-Sewda and the more extensive Wied il-Kbir system, also eroded parts of the Upper Coralline Limestone plateau away. Underneath the seaward boundaries of the plateaus lie the rdum lands (‘undercliffs’). These are near vertical rock faces formed either by tectonic movement or erosion (see below). Their bases are always surrounded by screes of boulders that eroded from the cliff edges (Giusti et al., 1995: 21).

Structurally, the Great (=Victoria Lines) Fault that crosses Malta in a north-easterly direction divides the island into two regions – a more fragmented northern part and a seemingly less strained southern part. This major structural feature is 14 km long and runs across the island from Ras ir-Raheb in the west to the Madliena tower on the east coast, the downthrow decreasing in height along the way from 180 m to around 90 m. (Vossmerbäumer, 1972:29). The scarp face is interrupted by several embayments, but only one makes a complete break through the Great Fault, albeit a very narrow one (Dewdney, 1961a: 34).16 North of this fault, a series of rift and horst blocks divide northern Malta into a landscape of valleys that are flat-floored basins, alternating with ridges as a result of tectonic activity (see above). These are, from the Victoria Lines Fault northwards: the Bingemma Basin, Wardija Ridge, Pwales Valley, Bajda Ridge, Mistra Valley, Mellieha Ridge, Ghadira Valley and Marfa Ridge. The sequence continues below the sea in the South Comino Channel, continues with the island of Comino as the next horst, followed by the North Comino Channel as the subsequent graben.

To the east of the Rabat-Dingli plateau and south of the Victoria Lines Fault, the landscape is characterised by a series of valleys that at times form deep depressions and were carved by streams, which descend to the central plains. Here, much of the Upper Coralline Limestone, Greensand and Blue Clay strata have been eroded away, thus exposing the Globigerina Limestone, and in the Attard and Naxxar areas even the Lower Coralline Limestone. Chalky and porous, the soft Globigerina Limestone weathers into layers with a low elevation, but as it contains several harder bands, these form steps in the landscape at each hard band. This results in the largescale gentle relief that leads to rolling plains and shallow depressions, are separated by low hills, which also characterise the surroundings of the Marsa Plain. The coring site inside the Marsa Sports Ground in Marsa is thus surrounded from east to south by fairly steep hills on which today we find Paola and Luqa,17 but the slopes become more gentle towards Zebbug and Qormi to the west whereas to the north there is the steep elevation towards Hamrun and the Sciberras Peninsula, where Floriana and Valletta are located today. At present, the Marsa plain nearly imperceptibly slopes down into one arm of the Grand Harbour towards north-east (Figure 2.7.). The hard bands within the Globigerina Limestone may also be impermeable and thus form minor spring lines through water percolation in areas where the hard bands meet the land surface (Pedley et al., 2002: 83).

South of the Victoria Lines Fault, the displacement is less conspicuous, as the following ridges and valleys converge like a fan in the wide alluvial Marsa plain: Naxxar-Gharghur Ridge, Lija-Msida Valley, AttardHamrun ridge, Wied is-Sewda, Zebbug Ridge, Wied ilBeqqiegha and Wied il-Hesri, Siggiewi Ridge and Wied Silani. These widien (singular wied) are water drainage channels that formed either by stream erosion during the much wetter Plio-Pleistocene, or by tectonic movements or a combination of the two (Giusti et al., 1995: 22; Pedley et al., 2002: 87). The topographic divisions south of the Victoria Lines Fault are clearly marked. In the west, two interconnected Upper Coralline Limestone plateaux form the most elevated area of the Maltese Islands, standing between 180 m and 245 m. This is also the most extensive outcrop of the Upper Coralline Limestone, and only here do all five strata still remain (Giusti et al., 1995: 21). It forms a continuous surface that covers around 23 km2 and consists of the Bingemma Plateau to the north and the Rabat-Dingli Plateau to its south (Nehring, 1966: 10). The area itself is shaped like a right-angled triangle that on all sides falls via steep slopes to lowland. Fretted all around by deeply incised valleys, the straightest and least eroded boundary runs between Ras ir-Raheb and Ghar Lapsi towards the sea, which is only interrupted once, by the regressing erosion caused by Wied l-Imtahleb and its tributaries (see Figure 1.2.). On its eastern boundary, the 16

In the southeastern part of Malta, a ridge that runs between Paola via Luqa towards Qrendi forms a watershed border that runs roughly northeast to southwest. Here, the widien drain in a southeasterly direction into the sea. Of these, Wied Dalam and Wied Has-Saptan are, in parts, deeply incised into the Lower Coralline Limestone and drain into Marsaxlokk Bay (see Figure 1.2.). The overall length of Malta’s coastline is around 137 km, which is roughly nine times its maximal southeast – northwest length. The southwestern part, between Benghisa Point in the southeast and Ras ir-Raheb in the west is formed by impressive steep cliffs with only few 17 Paola lies ca. 50 m a.s.l., Luqa 68m, Qormi 8m, Zebbug 108m, Mdina 200m, Hamrun 42m, while the Marsa Sports Ground lies on average 2.5m a.s.l.

This is the Wied tal-Isperanza.

19

Figure 2.7.: Steepness of slopes in Malta with Marsa catchment outlined and some localities. After Bowen-Jones, 1961.

incisions, where widien drain into the sea. From Ras irRaheb northwards, the coastline is characterised by semicircular and funnel-shaped shallow bays that are confined by promontories.

confined by low cliffs, of which Marsaxlokk Harbour is the largest bay (Figure 1.2.).

The coastline is similar between Ahrax Point, Mellieha, and St Paul’s Bay, but the northern coast between Marfa Point and Ahrax Point is characterised by flat karstland. The north- and southeastern coast consists generally of a subsided flat coastline, where the bays along its coast may often be regarded as a natural continuation of widien that were drowned. The drowned valley systems of Marsamxett and Grand Harbour are possibly the best example (Nehring, 1966: 13). Evidence, that these are drowned valleys includes, among other, the discovery of stalagmites on the seafloor when the breakwater in the Grand Harbour was constructed (Hyde, 1955: 103).

The main elements that control erosion and weathering are wave action on the coast, erosion by the sometimes swift-running streams during the rainy season, wind action, temperature difference between summer and winter and, also, during day and night in summer (Hyde, 1955: 17). Erosion, thus, manifests itself in various ways.

2.4.2. Rock erosion

Erosion of the rock is, among others, a result of often complex interactions between wind, sun and water: wind, for example, may easily pick up and carry siltsized particles, which cause friction on the exposed rock, when the dust brushes over it. When wind picks up moisture from the sea, the latter includes also soluble salts that then get absorbed by the rock, like a sponge, through capillary action. When the sun then dries the

Between Marsascala in the east and Benghisa Point in the south, the coastline is indented by shallow circular bays 20

stone, the heat generated on the surface forces the moisture out. By evaporation the salt crystallises inside the pores and pushes the pore walls out (Torpiano, 1996). Water, furthermore, reduces the strength of limestone considerably when tested in a saturated condition (ibid.). The limestones of the Maltese Islands are also dissolved by surface water and ground water. This is due to chemical reactions where rainwater becomes weakly acidic because it reacts chemically with carbon dioxide that occurs naturally in the atmosphere and the soil, forming carbonic acid. As the bedrock absorbs the water, the calcium carbonate, which makes up well over 90 % of the rock, gets dissolved along the surfaces of joints, fractures and bedding planes (Alpha et al., 1997: 3).

vary between less than 20 mm to more than 100 mm per thousand years (Grove & Rackham, 2001: 245). The dissolution creates hollows at joints in the rock and enlarges existing ones. The subaerial karst landscape of the outcropping Upper Coralline Limestone is particularly pronounced by the formation of narrow grooves with sometimes sharp edges (e.g. at Marfa Ridge). On Globigerina limestone, the resulting subaerial karst landscape gives rise, among others, to rolling hills and sinkholes and as well as to sinking streambeds. The latter are, however, more pronounced in the Lower Coralline Limestone (e.g. Wied Hanzira, parts of Wied il-Kbir and Wied Dalam). The absence of noteworthy perennial surface streams is a frequent feature in obvious karst landforms (Stafford et al., 2005: 15).

The various formations of the Tertiary geology that make up the Maltese Islands have a different resistance to the eroding elements, which also lead to the present landforms. Most prone to erosion is the Blue Clay as it is easily disaggregated and washed down by either rain or seawater. Wherever the Blue Clay is at or close to sea level, coastal landslips may occur as the sea erodes the clay and allows Upper Coralline Limestone slabs to collapse and slide downslope, as can be seen e.g. at Golden Bay in northwestern Malta.

The subsurface dissolution of the limestone by the slightly acidic running groundwater led also to the formation of large dissolutional closed depressions, some of which resulted in sinkholes or dolines once the ceiling caved in (e.g. Il-Maqluba in Qrendi), or caves, once a solution tunnel broke through to the surface (e.g. Ghar Dalam). In combination with soil, dissolution may also lead to the formation of small to medium sized shallow blind bulbous cave-like features called tafoni (Hunt, 1996: 57). More importantly, however, the saturation of the rock with rainwater and its storage within the rock provides a most important source of freshwater.

Globigerina Limestone is fine grained and only lightly cemented and thus easily worked as a building stone. However, although its surface may be ‘case-hardened’ by exposure to rain and sun (Pedley et al., 2002: 52-3), it is not immune to erosion and weathers either by flaking, dusting or honeycombing once its internal structure is weakened through continuous wetting and drying (Torpiano, ibid.).

2.4.4. Soils An important product of weathering and erosion are soils. Rocks that contain the highest amount of calcium carbonate yield a small residue after dissolution that consists of more or less insoluble clay, oxides of iron and manganese and other small mineral particles (Hyde, 1955: 77). Thus, the Maltese soils are characterised by their close similarity to the parent rock material from which they derive (Lang, 1961: 85). They are generally described as rather young or immature, because soil formation processes are slow in calcareous environments like the Maltese Islands, due to the very limited quantity of acidic drainage water that dissolves the limestone to form the soil (ibid.: 83). Consequently, time is an important factor in the soil’s development.

Both the Upper and Lower Coralline Limestone formations are more resistant to erosion than the Blue Clay and the Globigerina Limestone, but this is only relatively so. The Lower Coralline Limestone extends well below sea level where it is undercut slowly by wave action or, inland, is dissolved by groundwater in the aquifers. When exposed above sea level, it weathers, like the Upper Coralline Limestone under the influences of the sun, wind and rain, and in coastal areas, the sea. The erosive and weakening action of running water forms deeply incised valleys into the Lower Coralline Limestone, as for example parts of Wied il-Kbir in central Malta and Wied Dalam in the southeast. Large areas of the Upper Coralline Limestone, Greensand and Blue Clay have been eroded, thus exposing the Globigerina Limestone (see Figure 2.1.).

There are three main soil types in the Maltese Islands, all of which occur today within the Marsa watershed. The oldest of these soil types formed during the Pleistocene. Following Kubiena’s classification system (after Lang, 1961), these are the dark red to dark brown Terra soils, mature and extensively weathered, but little affected by the present climate because of the limited rainfall. In the Maltese Islands, the Terra soils are found on the karstland formations of the Upper and Lower Coralline Limestone and the Globigerina Limestone. Although all soils of the Maltese Islands have a low organic content, the Terra soils have the highest values with an average of 4.5%. The Alcol series may derive from Terra soils and develop on the valley loams. They are partly alluvial and partly

2.4.3. Karst features As nearly all the rocks of the Maltese Islands consist of limestone, there are characteristic morphological karst features. Karst is caused by the flow and percolation of rainwater during the wet season, but often the chemical solution of the rock plays a greater role than only mechanical erosion (Hyde, 1955: 20). The dissolution rates depend on the rock type and precipitation, but can 21

colluvial and occur in the broader valley bottoms (ibid.: 91-2, 88).18

Islands. The reasons for soil erosion are varied, and may involve climatic factors or anthropogenic factors or a combination of both. A crucial factor linked to erosion is the exposure of the soil. Soil depletion occurs when the soil formation is slower than its erosion. Considering the extremely slow soil formation process under the present climate (Lang, 1961: 83), its erosion may be considered a serious issue in the Maltese Islands.

Xerorendzinas have been developing over several hundred years only (ibid.: 83); they are immature, have a higher calcium carbonate content than Terra soils, but a lower organic content. Xerorendzinas chiefly develop on Globigerina Limestone and are brownish in colour, but also develop on valley loams (ibid.: 85).

Soil erosion may happen through wind, as here it may denude a rock from its overlying soil cover. On the other hand, soil particles carried by wind may form aeolian deposits, like at Wardija Point in Gozo (Figure 1.5., Hunt, 1997:106). The pattern of erosion and accretion of sand is strongly controlled by the vegetation cover and wind speed (Sarre, 1989: 17). In the absence of vegetation, the particles are easily picked up and transported away, while already very slight vegetation reduces significantly sheet and rill erosion (Grove & Rackham, 2001: 258).

Xerorendzina is the main soil type found in the Marsa plain today (ibid.: 84). The analysis of a 2 kg soil sample taken from the section of a ca. 60cm thick deposit in Marsa next to the ‘Millenia’ showroom in Albert Town (Grid Reference 545/704) revealed that this soil supported a light vegetation cover (presence of the molluscs Trochoidea spratti, Theba pisana and Cantareus aspersus).19 Carbonate Raw Soil is, according to Kubiena (in Lang, 1961: 87) ‘a very young raw soil, extremely low in humus, characterised generally by a dry summer (…) soil climate on calcareous parent rocks’. It develops on weathered calcareous Quaternary sandstones, the Greensand formation, on the lower beds of the Upper Coralline Limestone, on Blue Clay and on Globigerina Limestone (ibid.).

Both climate and topography affect the way in which erosion by water occurs. The high-velocity flows of steep slopes are more able to erode material than flow on gentle gradients, but it is not only the amount of rainwater available that is crucial. Very infrequent but violent rainstorms may result in far more erosion than greater amounts of water that fall steadily throughout the year (Nichols, 1999: 84). The presence of a vegetation cover may also be crucial in preventing erosion by runoff. Compared to an unvegetated or fallow surface, slopes covered by close-growing vegetation have been found to increase their resistance against erosion significantly, as vegetation protects the soil surface directly. Furthermore, the roots and rhizomes of the plants bind the soil and thus introduce extra cohesion over and above any intrinsic cohesion that the slope may have (Thorne, 1990: 135). The vegetation type may be an important factor as litter formation in a deciduous wood increases the degree of water absorption and thus favours runoff and fine sediment entrainment. A lower rate of litter and soil formation as found among evergreen oaks, on the other hand, makes the slopes more susceptible to debris displacement by gravity processes. Thus, the less a slope is protected, the higher the amount of sediment available to be eroded (Sala & Calvo, 1990: 347-8). A shrub matorral, even with a modest percentage cover, can provide erosion protection similar to that of trees, while a degraded matorral produces much higher rates of runoff (Francis & Thornes, 1990: 363). Simulated low-intensity rainfall events (around 25 mm/hr¯¹) on degraded and bush matorral have been shown to produce very little runoff on dry soil, but when the soil was wet, runoff was much higher (ibid.: 382).

Human agency also formed several soil complexes by mixing rock powder with already existing soil, by adding rock debris, and by adding domestic waste with the soil, and by mixing different soil types from various localities (ibid.). Several of these complexes may also be found within the Marsa watershed today (Lang, 1961: 84). The depth of the soils may vary greatly. It is rarely of a considerable depth and generally lies between a few centimetres and half a meter, depending on the topography. The greatest depths are found in alluvial plains (Nehring, 1966: 29), as is also the case for the Marsa plain (see below). Soils have been found to be deeper on the outcropping Greensands, Blue Clay and Globigerina Limestone strata than on the Upper and Lower Coralline Limestone, because soil forms more readily the higher the calcium carbonate content (Hyde, 1955: 77). However, it is the quality rather than the quantity of the soil, which is responsible for the fertility. Over the past two centuries, a large quantity of cultivable land has been prepared artificially by the farming population (Lang, 1961: 99), but despite the scarcity of soil, it is unlikely that any soil was ever imported in bulk by the Knights (Hyde, ibid.). 2.4.5. Soil erosion

Erosion is, thus, usually more severe during high intensity rainfall events. In the Maltese Islands, rainfall is predominantly convectional and often accompanied by strong to very strong winds. Mitchell & Dewdney (1961: 63) define a storm rainfall as when a continuous fall of at least 50 mm leads to significant runoff and physical damage, but add that these conditions are more likely to occur when rainfall exceeds 75 mm. These amounts may

Although soils are a product of weathering and erosion, they are also easily eroded themselves, which is another cause of the varying depth of the soil cover in the Maltese 18 They are also the erosion product of the more recent rendzinas and carbonate raw soils (ibid.). 19 The sample was taken by the author in February, 2000.

22

Figure 2.8.: Map of Malta by Jean Quintin (Quintinus), published in 1536 showing the Marsa Hortus and two possible drainage channels, the larger of which has a bridge, flowing into the Grand Harbour at Marsa. Shallow water and silts are indicated by dots in the inlets at Marsa and in Marsaxlokk Bay. The small building to the right of the bridge is probably Ghajn Filep, which supplied the ships with fresh water prior to the construction of the aquaeduct in the 17th century. Source: Agius-Vadala & Ganado, 1986.

overwhelm the absorbing capacity of the vegetation and soil. If vegetation and soil have been soaked by previous rain, nearly all will run off. When there is a minimum of plant cover, as may happen early in the rainy season, gullying and sheet erosion may ensue. Measurements generally show that disproportionally much erosion occurs during severe rainfall events (Grove & Rackham, 2001: 247).

1990: 139) and in connection with strong winds or storm, trees may easily get uprooted and may end up being carried away with the runoff, as appears to have happened in the Maltese Islands several times in the past, according to the accounts by Faurè (1913).20 Anthropogenic causes of soil erosion are varied: studies have shown that exposure of soil through overgrazing and poor land management can be a major source of rapid sedimentation (e.g. Meadows & Asmal, 1996: 42). Fires, whether wild or anthropogenic, may also lead to intensified soil erosion, because fires may cause the formation of a water repellent layer below the surface, which can inhibit normal infiltration of rainwater. As a result, the thus accelerated surface runoff has a highly erosive character (Boelhouwers et al., 1996: 9).

Soils derived from Blue Clay (e.g. the Fiddien series) are very heavy and thus not as easily eroded as terra soils or xerorendzinas on a comparable terrain (Lang, 1961: 99), but the steepness of the slope plays an important role, particularly in a natural environment with the absence of any artificial soil retaining measures like rubble walls and terracing. As the Marsa catchment extends all the way up to the Rabat-Dingli Plateau, the steepness of slope over which some of the runoff flows, at times exceeds even 40%, but for most of its path to the Marsa plain the gradient is between 5% and 15% (see above, Figure 2.7.). Thus, the topography influences erosion and particle transportation noticeably, while inviting the deposition of sediments and debris in the flat alluvial plain at Marsa. Quintinus’s 16th century map (Figure 2.8.) shows clearly the accumulation of sand and silt, where the streams flow into the sea in Marsa. These are likely to be the result of such erosive processes through heavy rainfall events.

Soil erosion may also happen in a less obvious way, which is, however, particularly favoured by the karst formations in the limestone and thus of importance in a Maltese context. As the insoluble residual minerals from limestone can form a soil together with windblown dust, they may also be washed down through fissures and disappear in caverns beneath (Grove & Rackham, 2001: 245). This may explain, for example, the presence of red soil in Ghar Dalam Cave near Birzebbugia, and also in the Hal Saflieni hypogeum in Paola.

Under former cooler climatic conditions soil crept slowly downslope as a result of solifluction, but under the present conditions, sheet or gully erosion may occur during severe showers (see above). This results in strikingly different eroded and alluvial phases of the same soil series in some locations (ibid.: 85).

2.5. Hydrology 2.5.1. Surface water – fluvial drainage The Maltese Islands also generally lack freshwater lakes (Nehring, 1966: 25 and see below); the two small freshwater bodies in Gozo21 qualify more as small ponds. Evidence of a past lake in (Middle?) Pleistocene Malta has been found at Fiddien Valley in the form of tufa

Trees are often credited as being the ideal prevention against erosion (e.g. Grech, 2001) and that intense erosion is due to lack of cover following deforestation (Francis & Thornes, 1990: 363). As mentioned above, experiments proved that shrubs are nearly as effective as trees, but furthermore, during very intense rainfalls, trees can actually aggravate the situation. Surcharge weight of trees may considerably reduce bank stability (Thorne,

20

Occasions when trees were uprooted and carried away by runoff were noted in 1343, 1530, 1755 and 1873, although this may have happened more frequently than recorded. 21 One freshwater body is located at Ta’Sarraflu, the other one is at Qawra (Figure 1.5.).

23

Figure 2.9.: Drainage patterns and valley systems of the Maltese Islands and their extensions below present day sea level. Source: Vossmerbäumer, 1972.

gorge in south Gozo may have had their origin when sea level stood close to the level of the present top edge of the wied. The beginning of the formation of the river valleys may thus date back as early as the Early Pleistocene, or possibly the Pliocene (Pedley et al., 2002: 87-8), while Vossmerbäumer (1972: 94) dates it back as far as the Upper Miocene. The stream would have needed relief energy to cut into the hard Lower Coralline Limestone, and this would have been controlled by the various cold stage regressions in the Pleistocene. By the Middle Pleistocene (Oxygen Isotope Stage 12) at the latest, when the sea level fell considerably, the relief energy for incisions into the Lower Coralline Limestone would have been sufficient (ibid.: 26).

deposits (Hunt, 1997 and see above), but the size and extent of this are unknown. The postulation of a former large freshwater lake at Marsaxlokk Bay (e.g. Keith, 1924) would need scientific evidence for confirmation, but is unlikely when considering the bathymetry (Vossmerbäumer, 1972: 101 and Figure 2.9.). Although several streambeds or widien may hold water for some time after rainfall events, notable stagnant and ephemeral lakes do not form.22 Waterflow, as well as transport of material by water is gravity-driven, in the sense that it provides the potential energy for the flow. As a result, the drainage pattern and the river types are controlled by the gradient, local vegetation and the proportions of the bedload and suspended load (Nichols, 1999: 83, 112). In the Maltese Islands, additionally to the streambed erosion, many riverbeds or widien are also controlled by the faulting (Vossmerbäumer, 1972: 81) or a combination of the two processes (Schembri, 1993: 30). As, generally, rivers tend to meander at gradients lower than 0.1º (Nichols, 1999: 112), snake-like courses like that of the Wied Hanzira

Apart from the above-described meandering type, the majority of widien exhibit a tributary drainage pattern, which is often one-sided (e.g. at Mellieha and Marfa Ridge, and also the Wied il-Ghasel has tributaries only on one side). The Marsa system, which is also the most extensive drainage system on Malta, exhibits a roughly dendritic pattern for Wied il-Kbir, but a more one-sided tributary pattern for Wied is-Sewda (see Figure 2.10.). The tilt of the Maltese Islands to the north-east ensures that most of the surface water is directed in this direction due to gravity. As a result, many valleys are roughly

22

The ‘Chadwick Lakes’ are artificial reservoirs behind dams that serve to retain the floodwater after rainfalls to allow percolation and infiltration through the bedrock into the Lower Water Table (Giusti et al. 1995: 24)

24

Figure 2.10.: Physiography of Malta with main river valleys and Marsa catchment outlined. After Dewdney, 1961.

aligned along or parallel to the SW-NE faults and the position of the fault lines on the Maltese Islands strongly controls most of the valleys (Pedley et al., 2002: 86). The directions of the present watersheds indicate changes in the drainage patterns such as run-off changes within valley segments. Indication of this may be seen, for example, at Wied il-Ghasel, whose former possible drainage at Qalet Marku is very suggestive (Vossmerbäumer, 1972: 86). Furthermore, in northern Malta, a N- to NNW oriented valley system would have been interrupted by younger tectonic movements, thus desiccating valleys on the edge that were suddenly disconnected from the hinterland (ibid.: 90). Therefore, today’s drainage system is unlikely to be the original one. More recent tectonic activity is suggested by hanging valleys (e.g. at Paradise Bay, see above), but also by drowned valleys, as indicated by the submersion along the eastern- and southeastern coast. The embayments (creeks) within the Grand Harbour are good examples of drowned river valleys (Pedley et al., 2002: 94) and a deep erosion of the valley floors is likely to have been promoted by lowered sea levels in various glacial stages

during the Pleistocene (Vossmerbäumer, 1972: 92). A continuation of these now drowned river valleys can be traced by the depth sounding of published marine charts (Figure 2.9. or less detailed Pedley et al., 2002: Fig. 122). A problem here is that these marine charts give the depth of the sea floor, but not of the bedrock, which may lie buried under a considerable layer of sediments. 2.5.2. Surface water in the Marsa catchment The large amounts of run-off water carried by the widien enabled their streambeds to incise themselves at times deeply into the prevalent Lower Globigerina Limestone. The depth may be indicative of significant rainfall periods or events during various past Pleistocene regressions (Hyde, 1955: 18). A clearer idea of the strong erosional force of water carving its way through the Lower Globigerina Limestone may be deduced from the topographical survey of the bedrock sea floor of the streambed (i.e. beneath the sediments), into which the dock for Malta Shipbuilding had been constructed (see below, Figure 4.6. and compare Figure 2.11.): over a 25

distance of less than 500m, the valley floor drops more than 15m. Within this streambed, several large potholes can be identified, which were possibly formed due to the force of the water (Hyde, 1955: 18). The water catchments of the tributaries that drain into this streambed collectively do not exceed 5 km2, the subaerial slope that extends from Marsa up to Luqa does not exceed 5%, although the comparatively short slopes from Paola Hill (between 5% and 15%) may contribute to the speed with which the water would erode the streambed (Lang, 1961: 98). Although the main tributary that drains into the streambed is marked out as a wied on older geological maps (e.g. in Murray, 1890; Dewdney, 1961a: 35), this wied no longer features as such in more recent maps. Abela’s 1647 description does not mention it either. The considerable erosional depths achieved by this rather inconspicuous and small-scale slope system provide the suspicion that the true depth of the known drowned river channels may be even deeper. Transposing this erosional force of water from the streambed at Malta Shipbuilding onto the apparent plain of the Marsa Sports Ground suggests a possibly deeply eroded area of a considerable size as here the much more extensive drainage pattern of both Wied is-Sewda and Wied il-Kbir and of several unnamed tributaries shown on the topographical map of the Marsa basin (see Figure 2.11.) drain into the area now occupied by the Marsa Sports Ground. During the various Pleistocene regressions, the gravitational force of the flowing water would have been much stronger here due to the lowered sea levels. With the Holocene transgression, the possibly very wide river valley would have been drowned by the rising sea level.

The present surface topography outlined on Figure 2.11 may perhaps support this hypothesis. Today, at the Marsa Sports Ground no visible trace remains of any other sort of stream, seasonal or perennial, nor of any brackish water body that would have been formed at the mouth of the streams and their tributaries, where they flowed into the sea. Blouet (1964a: 200) mentions the presence of two lakes in the 17th century in Marsa, although ‘pools’ may perhaps be a better term. Presently, even within the whole Marsa catchment, there is no perennial freshwater body. 2.5.3. Subsurface water It would be misleading to deduce from the apparent scarcity of surface water a similarly deficient ground water reserve (Nehring, 1966: 25). Although the natural water resources of the Maltese Islands depend entirely on rainwater (Schembri, 1993: 31), the geological stratigraphy (see above) of the Maltese Islands provides particularly favourable conditions for the percolation and underground storage of rainwater because of the porosity of the limestone (Dewdney, 1961b: 43). Rainfall is highly variable from year to year: for the period between 18541990, the minimum recorded was 191.3 mm, while the maximum was 1031.2 mm (Chetcuti et al., 1992). The recent average annual rainfall is around 517 mm, which translates into 8,900,000 m³ (mean between 1951-1990), of which 6-20% are run-off, up to 70% are lost by evapotranspiration and the remainder (up to around 24%) recharges the aquifers (Planning Authority, 1990: Q20;

Figure 2.11.: Topographical map showing a detailed drainage pattern of the Marsa catchment with floodplain. Grid lines follow north-south and est-west directions, each square equals 1 km². Source of base map: Malta Environment and Planning Authority (MEPA), 2003.

26

Mangion, personal communication: 200023). These figures vary, among others, because runoff depends on the saturation of the soil and permeability of the surface, and evapotranspiration on temperature, wind and humidity. Nonetheless, these figures highlight that much of the drainage occurs underground rather than as surface runoff. This is typical for karst areas (Alpha et al., 1997: 4-5).

rapid run-off (Dewdney, 1961b: 43). The Mean Sea-level Aquifer is the largest aquifer and has been a major source of the public supply since the mid 19th century, when it was exploited by cutting shafts and galleries into the rock. It occupies an area of 107 km² (ibid.) and consists of a freshwater lens floating on the denser saline water at sea level in the limestone rock. Perched aquifers are the other important aquifers. Here, the rainwater is trapped in the permeable Upper Coralline Limestone by the underlying impermeable Blue Clay. This is the case mainly in the western part of Malta and on the eastern and central parts of Gozo. Wherever the Upper Coralline Limestone/Blue Clay interface is exposed, water can seep from the Perched Aquifers to form so-called High Level springs, which then drain into the widien watercourses (Giusti et al., 1995: 24). Several of these used to flow all year round, even though with a much reduced flow

The Upper and Lower Coralline Limestones support the most important aquifers, but the presence of bedding planes, joints and fissures controls the underground movement of water and is, as such, more important than the porosity and the permeability of the rocks themselves. The Globigerina has an overall low porosity and percolation is also slow because of the presence of thick and relatively impervious marly beds. This results in a

Figure 2.12.: Hydrology of Malta after T.O. Morris, 1952, with Marsa catchment and Marsa Sports Ground marked out. After Dewdney, 1961. 23

Dr John Mangion, Water Services Corporation, Luqa, Malta.

27

during the dry period. However, most of these are now tapped by farmers, which results in predominantly dry valley beds. While Quintinus in 1536 points out that the freshwater tastes salty and sedimentary and emphasizes its dependence on the rainfall (in Vella, 1980: 38), Abela (1647: 92-3) praises its abundance, sweetness and exceptional quality. Perhaps more objectively, scarcity or availability of freshwater appears, in historical times at least, to have become a serious problem only as recently as the 19th century due to rapid growth of the population and a rising standard of hygiene and living (Hyde, 1955: 94; Dewdney, 1961b: 43).

form one landmass out of all Maltese Islands (Vossmerbäumer, 1972: 97), and a fall of 155m would create a land bridge with Sicily (Hunt & Schembri, 1999: 67). On the other hand, a 10m rise of the present sea level would submerge large parts of the low-lying areas along the north-eastern and south-eastern coast, flooding Marsa, Sliema and Marsaxlokk, amongst others. The sea surface generally rises and falls with tides and waves, and with changes in the atmospheric pressure, wind, temperature and salinity (Pirazzoli, 1998: 5). Presently, there is relatively little variation in the sea level around the Maltese Islands. The maximum tidal range does not exceed 20cm (Drago, 1993), although apart from short-lived weather driven sea surges of varying heights, the northern coast of Malta is also subjected to non-tidal short period sea level fluctuations known by local fishers as ‘il-milghuba’. Of atmospheric origin (these phenomena develop as a result of storms or of rapid changes in air pressure), these coastal seiches consist of sea-level oscillations with periods ranging from a few minutes to several hours, often accompanied by remarkable currents that may cause severe damage to coastal areas (Drago, 2000).

2.5.4. Underground water resources in the Marsa catchment The Mean Sea Level aquifer presently underlies a large area of the Marsa catchment, but exploitation by the use of boreholes only reaches down to the limits of Qormi and thus does not extend into the Marsa Sports Ground (see Figure 2.12.). Abela (1647: 92-3) mentions many gardens and horti that used to thrive along the valley plain from Qormi towards Marsa because of the fresh water that then apparently was available in large quantities, undulating everywhere along the plain and providing ample supply for all gardens and horti. He further documents the presence of four large freshwater wells.24 Aiyn Filep, also prominently shown on several old maps (see e.g. Figure 2.8.) was a place close to the sea from where ships used to replenish their supplies, until water became available in Valletta. The water from Aiyn Filep was famed for its excellent keeping qualities, which even on long trips remained uncontaminated (Abela, ibid.). Aiyn Filep was probably a water seepage from the Globigerina in the form of a spring around which a stone structure was constructed so that the spring flowed through jets and into troughs or a large bowl (Schembri, personal communication, 2006). Today, the place no longer exists. The whereabouts of the other wells is also uncertain, as some of them may have been lost beneath the expanding periphery of Qormi, while others could have lost their old name (Galea, 1961: 44). Due to the mixing of freshwater with the underlying seawater, the limit of the extractable potable water from the Mean Sea Level Aquifer reaches down to Qormi, and it is possible that the old wells mentioned by Abela, already drew their freshwater from this aquifer.

When these periodical and random movements are filtered out, the mean sea level (MSL) is obtained (Pirazzoli, 1998: 5). In the Maltese Islands, sea levels started being measured in 1876, when a gauge was installed in French Creek (near Cospicua) within the Grand Harbour, but there are no measurements after 1926. Since 1988, the Malta Maritime Authority has operated a sea level gauge in the Grand Harbour, and established the mean sea level through analysis of 13 months of data,25 while at present sea levels are also being measured at Portomaso (St. Julian’s) and at Mellieha (Drago, 2000). 2.6.1 Eustatic (global vertical) changes While the above are comparatively minor changes in the present sea level, much more dramatic changes occurred in the past. Global climate changed rapidly and constantly in the Quaternary, predominantly due to the Milankovich factors of orbital eccentricity (100 ka periods), with numerous superimposed glacial advances caused by orbital obliquity (41 ka periods) and precession cycles (23 ka periods). This resulted in alternations between glacial cold periods (stadials) and temperate conditions similar to today (Hunt & Schembri, 1999: 41-2). These alternations are well expressed for the past 2.6 million years, and are deduced from oxygen isotope records obtained by analysing the benthonic foraminifera of the deep-sea sediment core ODP core 677 (see Figure 2.13.) and shows that climate change is indeed the main cause of changes in the quantity of oceanic water (Pirazzoli, 1998: 7).

2.6. Sea level and coastal changes since the Pleistocene A significant rise or fall of sea level has far reaching consequences for the landscape and the hydrology of the island. A fall of ca. 9m would considerably decrease the present size of the harbours and extend the coastline along the north-eastern coast (see Figure 2.9). As a result of the lower sea level, the Mean Sea Level Aquifer, would also be lowered. A sea level fall of ca. 24m would 24 Bir Eebeyer, Bir el Meru, Bir Buhagiar, Bir el Vasa. Of these, Bir Buhagiar is said to contain perennial running fresh water in large quantities.

25 The data obtained between May 1990 and May 1991 was used to establish the tidal harmonic constants.

28

interrupted land bridge between Malta and Sicily during at least some of the past glacial periods. A land bridge in the late Early Pleistocene may have led to the earliest terrestrial remains of vertebrates on Malta (Leithia carthei stage at Ghar Dalam; Storch, 1974: 430). Earlier remains of vertebrates have, so far, not been found. Considering, however, that a preceding transgression of the Mediterranean may have reduced the size of Malta possibly to a series of smaller islands formed of the present day Upper Coralline Limestone Plateaux26 (Vossmerbäumer, 1972: 24), this would be hardly surprising. On the other hand, a connection with Tripoli in North Africa would require a lowering of around 400m and still leave one or two deep channels (Hyde, 1955: 28). Thus, a connection with the African shelf is unlikely to have occurred during the Pleistocene (Pirazzoli, 1998: 10). Changes in the sea levels in the Mediterranean since the LGM have been the focal point of many studies (e.g. Lambeck & Bard, 2000; Antonioli et al., 2003; Lambeck et al., 2004), but sometimes the accuracy of sea level curves derived from the data is debatable (see Pirazzoli, 1992: 21-2). Isostatically, the data have been found to vary regionally as well as locally, at times considerably even over short distances (e.g. Antonioli et al., 2003 for Sicily), which may lead to inaccuracies when transposing data from other regions in the Mediterranean to the Maltese Islands. However, eustatically a general trend may be observed: By the last glacial maximum (LGM) around 20,000 to 18,000 years ago, it is estimated that the world’s oceanic sea-level had fallen considerably. As global changes indirectly also affect the Mediterranean region, it is presumed that large areas of the central Mediterranean where the present sea floor is less than 100 m deep would have been dry land during the LGM (Pedley et al., 2002: 90) and other glacial periods of a similar character. The postulated maximal sea level fall in the Mediterranean before the last transgression varies from around 150 m at 20,000 BP in parts of the central Mediterranean (Lambeck et al., 2004: 1594) to 120 m in South France according to Lambeck & Bard (2000) to around only 90 m generally according to Vossmerbäumer (1972: 25). For the Maltese Islands, a remarkably even wave-cut terrace below the 50 fathom (fm) line (ca. 91 m) may correspond to a possible lowstand of the last regression (ibid.: 98). According to Pedley et al. (ibid.), a land bridge that connects the Maltese Islands with Sicily would require a lowering of sea level of some 100 m, but this appears to be an idealised picture since at present, such a regression would leave a more or less 14 km wide channel between Malta and Sicily (Hunt & Schembri, 1999: 67). In fact, at present, a sea-level level drop of some 155 m is currently needed to connect the two islands together, although it is

Figure 2.13.: Oxygen isotope record for the past 2.6 million years deduced from benthonic foraminifera of the ODP core 677 with labels of selected isotope stages added for orientation. Odd numbers correspond to interglacials, even numbers to glacial stages. Source: Pirazzoli, 1998.

During the last 2.6 million years, there were at least 100 major changes between cold and warm stages, with many smaller ones superimposed. Stages 12 and 16 correspond to exceptionally large glaciations, where the ice volume is estimated to have been around 15% greater than at the last glacial maximum (Stage 2). The warmest interglacials with a higher global sea level occurred at Stages 1, 5e, 9 and 11, but it cannot be stated confidently whether the sea levels attained in any one of the stages differed significantly from any other, only that as a result, with the development or melting of continental ice sheets, sea levels changed accordingly (Pirazzoli, 1998: 9). During glacial periods, the water that evaporated from the sea fell as snow on land and became ice in great continental ice-sheets in Northern Europe and North America. As a consequence, the sea level fell considerably (Hunt & Schembri, 1999: 44). There are strong indications that there may have been a complete or

26 The Calabrium transgression is said to have caused sea levels to rise globally by around 180 m above present day level, while the subsequent Sicilium transgression possibly raised sea levels between 80 m and 100 m above present day levels (Vossmerbäumer, 1972: 24).

29

highly probable that in the past, tectonic movements could have considerably altered the submarine topography (Vossmerbäumer, 1972: 109).

led additionally to a subsiding sea bottom (hydroisostasy) (Pirazzoli, 1998: 13). Archaeological remains may be useful in helping to establish former sea levels (e.g. Blackman, 2005: 61; Auriemma et al., 2005: 5), but there are limitations as most archaeological remains give no evidence on how far from sea level they were originally located. Harbour constructions, such as slipways and especially fish-tanks may provide precise indications, but unfortunately, in the Maltese Islands, these have either not been found or have meanwhile been destroyed.28 Vossmerbäumer (1972: 8) mentions steps dated to the Roman Period on the east coast of Cominotto, which end considerably above present day sea level. While the steps may be evidence of either uplifting or a formerly higher sea level, it is, however, not possible to confidently date these stone steps to any period. Undated rock terraces above the water line on the west coast of Malta (around Gebel Ciantar) and in Xlendi Bay (Gozo) may also be evidence of formerly higher sea levels and/or uplifting (ibid.). Several other now submerged archaeological features (e.g. the Bronze Age silo pits at St. George’s Bay, Birzebbuga, as well as perhaps the cart ruts at the same locality) may point to lower sea levels during the Bronze Age, but may equally indicate subsiding of the land mass since the Bronze Age. Thus, although the local relative sea level has risen, it is impossible to ascertain without further scientific investigations whether this is due to a eustatic rise or an isostatic phenomenon or a combination of the two.

During the transition from the last glacial period to the Holocene, global temperatures rose rapidly, as did the sea level (Hunt & Schembri, 1999: 44). In Marseille in South France, sea levels have been found to rise regularly up to 500 AD, followed by a period of stability. Since 4,400 BP, the total rise has here been less than 1.5 m (Morhange, 2005: 26). In Mykonos, Delos and Rhenia in the Eastern Mediterranean, however, relative sea levels lay uniformly 3.8 m below present day sea level in 1800 BC, -2.5m 2000 years ago and around –1 m in 1000 AD (Fouache et al., 2005: 37). The coastal areas of southern Apulia (Italy) show a different picture. Here, studies show that between Bari and Taranto, sea level was at least 1 m above present level during the maximum Holocene transgression around 7,500-7,000 BP, and fell 3 m below present level during the Bronze Age to rise slowly afterwards to the present level (Auriemma et al., 2004: 19). Similarly, between Carbinia and Egnatia in Apulia, sea level was at least 1 m above the present one around 6,000 BP, to slump 1.5 m below the present position by 2,500 BP followed by a slow sea level rise up to the present position (Mastronuzzi & Sansò, 2002: 139). Also near Brindisi in Apulia, there are indications that sea level at 3,300 BP was 3 m below present level and was raised to 2.5 m lower than present by 2200 years ago. The following rise to present level then flooded the medieval structures (Auriemma et al., 2005: 5). These studies may confirm that in the last 5000 years, global sea level has been dominated also by dynamic climatic-oceanographic variables (Mörner, 2005: 99).

Generally, for the Maltese Islands, the sequence and timing of the stages of the rising sea levels since the LGM have as yet not been specifically studied. However, a lot of research has been carried out to investigate tectonic movements and relative sea level changes along the Italian and Sicilian coasts. Data from the tectonically active Sicilian east coast indicate that the coast at Taormina has experienced an uplift of ca. 8.4m in the last 6000 years, but that the southern coast has remained largely unaffected by tectonics (Antonioli et al., 2003: 1). Relative sea-level change along the Italian coast and adjacent seas has been found to exhibit considerable spatial and temporal variability throughout the Holocene (Lambeck et al., 2004: 1567). Combining eustasy, glaciohydro-isostasy and vertical tectonic motion, Lambeck et al. (2004) remodelled the central Mediterranean coastline for the past 20,000 years. Extrapolating the data from the research of Lambeck et al. (2004), the sea level of the Maltese Islands (relative to the present MSL) during the past period may have been as follows: • 20 000 BP = -150m • 16 000 BP = -120m • 12 000 BP = -75m

2.6.2. Isostatic changes in the sea level While eustatic sea levels may give a low resolution picture of global changes, it has been found that sea level changes have varied greatly from place to place since the last glacial maximum (LGM), particularly as more data become available (Pirazzoli, 1998: 3). Thus, apart from global and regional eustasy, isostasy may play a major role. Isostatic variations in the sea level may be due to a variety of reasons, but in the Maltese context, tectonic uplifting and subsidence may be the most important ones,27 as a result of earthquakes but also as a result of tilting. Little is known about past effects of tectonics on the relative sea levels and also of the ongoing tilting, neither the uplift rate along the north-western coast, nor the submerging rate along the north-eastern coast have been the subject of a scientific study (Pedley et al., 2002: 90). What is certain is that with rapidly rising sea levels at the onset of the Holocene, the formerly much more extensive coastline and valley bottoms became submerged (see Figure 2.9.). As a result, the meltwater from the ice sheets produced a considerable load (estimated 1t m¯² per 1m sea level rise), which may have

28 E.g. a pile of timber, possibly a pier was found under a thick layer of clay at roughly present day sea level during the construction of the dock for Malta Shipbuilding Co Ltd. It may have been a pier during the Punic/Roman Period, which might indicate that formerly the sea level had been lower, as a pier is generally higher than sea level. However, all datable evidence is lost (Saviour Scerri, geologist at the construction site of Malta Shipbuilding Co Ltd, personal communication, 2000).

27

Erosion may also affect the relative sea level in places through the deposition or removal of large amounts of sediments.

30

• • • • •

10 000 BP = -50m 8 000 BP = -21m 6 000 BP = -9m 4 000 BP = -4.75m 2 000 BP = -1.75m

Although the data of Lambeck et al. (2004) do not include the high sea level stands registered in Apulia by Auriemma et al. (2005: 5) and Mastronuzzi & Sansò (2002: 139), in the absence of any other or more precise scientific data, the results of Lambeck’s et al. (2004) study will be used in the present study to provide a preliminary background for the environment and the archaeology and, where possible, will be tested against the data from Marsa Core 1. What already emerges from the data of Lambeck et al. (2004), is that by the time the first settlers would have arrived in Malta around 7000 BP, the Maltese Islands would have probably already been split up into Malta, Gozo and Comino, as the current minimum depth that separates Malta from Gozo is 24m (Vossmerbäumer, 1972: 97).

31

Chapter III

of Gozo, Ltd. John Otto Bayer, excavated at the Xaghra Circle and cleared nearby Ggantija Temples. In 1839, the temple of Hagar Qim was cleared by Col. J.G. Vance and this was followed by the temple of Mnajdra in the following year.4 Comparing, however, Houel’s drawing of Hagar Qim (1787) with the plan and illustrations made by Vance’s draughtsman Foulis following the excavation, it has been suggested that Hagar Qim may have been reconstructed prior to Vance’s excavation in 1839 (Mayerhofer, 1996). Although Vance’s excavation report (1842) with drawings by Foulis may have its shortcomings and has often been criticised (e.g. Mayr, 1909; Evans, 1971), it provides valuable information. Much less is known of the clearance of Mnajdra (Trump, 2002: 7). Of Bayer’s sites, the drawings and watercolours of Charles Brochtorff provide an important basic record (Trump, ibid.). After that, the antiquities of the Maltese Islands did not receive much attention or care, and in 1871, Cesare Vassallo, librarian and curator of the many artefacts discovered so far, complains bitterly about the bad state of preservation of, among other sites, St. Paul’s Catacombs.5 After a complaint reached the House of Commons saying that the antiquities of Malta were in a shabby state and were not being catered for properly, in 1881 the Governor of Malta instigated A. A. Caruana, then Director of Education, to prepare a Report on the Phoenician and Roman Antiquities on the Islands of Malta, which was published in 1882 (Mizzi, 1982: 874). This report revived interest in the islands’ antiquities. Caruana’s main interest lay with the palaeochristian and pagan past, in the study of which he is considered to be a pioneer (e.g. Heritage Malta website6), although he has also often been criticised for his uncritical and negligent approach (e.g. Mayr, 1901: 216 and Becker, 1913: 5).

ARCHAEOLOGICAL BACKGROUND 3.1. Early descriptions of monuments and history of archaeological research in the Maltese Islands The Maltese Islands occupy only a small space within the central Mediterranean, but they boast an enormous wealth of archaeological remains from the Neolithic Period onwards. Ptolemy and Cicero provide what are possibly the earliest references to monuments, but the temples both mentioned were contemporary to the Classical Period in which both authors lived, and are thus not mentioned in an archaeological context. The looting of one of these temples supplied fuel for Cicero in his orations against Verres,1 while the location of two temples2 provided points of reference in Ptolemy’s Geography. No further mention of monuments is made in literary sources until the 16th century, when the French Knight Jean Quintin d’Autun (Quintinus), after having spent three years on the islands, published his description of Malta in 1536. Quintin d’Autin mentions vast ruins at Marsaxlokk Bay and links these with Ptolemy’s temple of Hercules (Vella, 1980: 23), but whether Quintin d’Autin refers to the prehistoric remains in Borg in-Nadur (according to e.g. Vella, 2002: 85; Bonanno, 1982) or to the sanctuary at Tas-Silg (according to e.g. Trump, 2002: 6) is uncertain. Gian Francesco Abela (1647) described and documented many more archaeological remains and, recognising their antiquity, housed an important collection of artefacts in his Villa San Giacomo, which was located on today’s Jesuits Hill in Marsa. Soon out of print, his extensive publication was republished by Count Ciantar in 1772, who added some archaeological sites that had meanwhile been discovered.3 Among these were structural remains that came to light during works in the Grand Harbour in 1768 (Abela & Ciantar, 1772: Lib I, Not. III). Dated to the Roman period, their discovery was also documented (and their destruction condemned) by Marquis Barbaro, whose observations were eventually published in 1794. As international interest in the antiquities of the Mediterranean grew, Malta often formed part of the itinerary of travelling antiquarians or engravers. Here, Jean Houel stands out. The 1787 publication of his Voyage pittoresque des îles de Sicile, de Malte et de Lipari contains, apart from descriptions, also plans and views that show how several sites appeared more than 200 years ago.

The Jesuit Fr Manuel Magri was assigned to excavate several sites in Gozo from 1889 onwards, and after Caruana’s death he was entrusted to excavate the then recently discovered hypogeum at Hal Saflieni in Paola until he was sent to Tunisia, where he died in 1907. Of his major excavations, only the report of the Xewkija remains has survived. Unfortunately, of the Hal Saflieni hypogeum Magri’s reports or notes have not been found (Trump, 2002: 7), and information about the excavation found in letters is scanty (Briffa, 2002-2003: 41-46). Meanwhile, an interest in Maltese antiquities was also generated in Germany and here, Albert Mayr, a philologist, contributed greatly to the knowledge of Malta’s past. Among other achievements, he recognised the prehistoric character of the temples, and convincingly argued against a Phoenician origin (Stoeger, 2000: 6). In 1904, Themistocles Zammit was appointed Director of the Malta Museum, and with him, excavation techniques and recording began to make a big leap forward. What

The first documented clearances of prehistoric remains began in the 1820s in Gozo, when the then Commissioner

4

Both temples are located south of Qrendi in the south of Malta. “E qui mi si permette un lamento. – In quale stato di conservazioni versano le Catacombe di S. Paolo. . . Non tanto le condizioni dei luoghi ed il volgare dei secoli, quanto l’imperdonabile incuria di coloro cui corre il debito di prenderne pensiero, farà si che fra non molto ce ne rammenteremo come di cose che già furono.” (Vassallo, 1871: 23). 6 http://www.heritagemalta.org/catacombs.html as on April, 2006.

1

5

The Temple of Juno, now identified with the sanctuary at Tas-Silg in Marsaxlokk 2 Ptolemy mentions the Temple of Juno, and the Temple of Hercules, possibly located at Ras ir-Raheb (Vella, 2002: 83 and below). 3 For example the Roman burial site that had been discovered near the Marsa at Qormi.

32

followed was a burst of activity, led by Zammit himself, Thomas Ashby and T.E. Peet, the German Erich Becker and many others. Zammit continued the excavations at the Hal Saflieni hypogeum and subsequently published his findings, but an archaeological breakthrough came with the excavation of the newly discovered prehistoric temple at Tarxien in 1915. Here, a clear stratigraphy finally allowed a division of the pottery found into Neolithic and Bronze Age. After World War I he excavated several other temples, notably Mgarr between 1923-7 (MAR, 1927). His meticulous and careful approach earned him deep respect and admiration also beyond the shores of the Maltese Islands (e.g. Becker, 1913: ii), while his publication of detailed excavation reports, apart from discoveries mentioned in the Museum Annual Reports, ensured that a written record remained along with the properly stored material record.

publicised Malta’s archaeological wealth. Seen as a hindrance to fast economic development, many archaeological remains sadly never made it to the surface during the building boom in the 1970s.7 Adding to that, local archaeology took a setback with the amendment to the Education Act in 1980, which suppressed the Faculties of Arts and Science until they were reinstated in September 1987. Since then, the Department of Classics and Archaeology of the University of Malta embarked on a joint excavation with a number of British universities at the Xaghra Circle in Gozo between 1987 and 1994, which revealed marvellous artefacts and helped to add considerably to the knowledge of the Temple Period. Between 1996 and 2004, parts of the southern area of Tas-Silg were excavated by the Department of Classics and Archaeology at the University of Malta. These excavation seasons also served to provide excellent training opportunities for future archaeologists.

After Zammit’s death in 1935, the curator of the Natural History Section within the Museum Department, J.G. Baldacchino was appointed Director of Museums. It may well be thanks to his scientific background that under his directorship more attention was given during excavations to accompanying molluscs and to soil texture (see e.g. the report of the discovery of the Zebbug tombs, MAR 194748: I-II).

3.2 Chronology The groundbreaking work by both Evans and Trump resulted in a chronological order of the cultural sequence in the Maltese Islands based on pottery typology from secure stratigraphic contexts. While the established sequence enabled relative dating, eventually absolute dating in the form of radiocarbon dates allowed points in that sequence to be fixed on a time scale.

In 1952, John D. Evans was placed in charge of the Survey of Prehistoric Monuments of Malta by the Royal University of Malta, and he produced a relative chronology from the vast amounts of prehistoric pottery stored at the Museum Department as well as an integration of the development of the temples into his scheme. His 1959 and 1971 publications provided a sound base for later studies on Maltese prehistory, even if in the absence of absolute dating methods, his assigned dates were later found to be too low (Trump, 2002: 8).

3.2.1 Radiocarbon dating Radiocarbon dating has been one of the most significant discoveries of the 20th century science as it allows to establish a chronological sequence and put a date on it (Higham, 1999). Called by Renfrew the “Radiocarbon Revolution” (1973), the method also revolutionised Maltese Prehistory as it revealed an antiquity of the prehistoric monuments considerably older than previously thought.

David Trump’s appointment as Curator of Archaeology at the National Museum between 1958 and 1963 led to further excavations at Borg in-Nadur, Bahrija and, most importantly, at Skorba. Here, the entire Neolithic sequence was found in stratigraphic order, while two new phases were discovered, as well as traces of a settlement site. With the development of the radiocarbon dating method, the date of the arrival of the first settlers could be established, as well as the dating of the Temple Period and the beginning of the Bronze Age (Trump, 1966 and see below). The discovery of Bronze Age cisterns at TalMejtin near Luqa provided the material for the first pollen analyses (Godwin, 1961: 8).

The radiocarbon method is based on the decay of the naturally occurring but highly unstable C14 isotope, which is formed in the upper atmosphere through the effect of cosmic ray neutrons upon nitrogen 14 (N14). Oxidised to 14CO2, C14 is absorbed through photosynthesis and the food chain by plants and animals during their lifetimes. As soon as a plant or animal dies, the carbon uptake ceases and the radioactive carbon present in the tissues starts to decay at a steady and accurately predictable rate. In this decay process, Libby (1955) found that the number of radioactive carbon atoms is halved after 5568 years, but the figure has meanwhile been revised to 5730 years (Godwin, 1962: 944); therefore, the age of death of the plant or animal tissue can be calculated by measuring the amount of

Major excavation campaigns also took place in the 1960s by the Italian Missione Archeologica at the Roman villa at San Pawl Milqi in Burmarrad and at the Tas-Silg sanctuary in Marsaxlokk. The 1970s, excavations at Hal Millieri near Mqabba helped to shed more light on the medieval period (Blagg et al., 1990), while Colin Renfrew’s chiefdom model of the Maltese Temple Period in his vastly popular Before Civilization (1973), together with an article about the Maltese Temples that appeared in 1977 in the National Geographic magazine, widely

7 For example, archaeological remains discovered during the excavation of a dock for Malta Shipbuilding disappeared or were destroyed according to Saviour Scerri, who was then employed as a geologist at the construction site (see below), or a vast number of Punic tombs that vanished during the construction of the Government Housing Estate at Tarxien near the temples, according to many residents who now live there.

33

radiocarbon left in a sample. The accurate measurement of the C14 content of a sample is subject to counting errors, background cosmic radiation and other elements that add uncertainty to the measurements. This probable error is expressed in the standard deviation (e.g. +/- 40) attached to every radiocarbon date (Renfrew & Bahn, 1996: 132). This measurement error represents the standard deviation (1σ) on either side of the quoted date (that is, 68.2% probability distributed symmetrically either side of the quoted age, which is the centre of the Gaussian probability distribution). A 2σ standard deviation is exactly double the 1 standard deviation (i.e. here +/- 80) and gives a 95.4% confidence region. The conventional radiocarbon age is given in radiocarbon years before present (BP), where ‘the present’ is conventionally fixed at 1950.

3.2.2. The Maltese Radiocarbon dates So far, there are 31 radiocarbon dates published for the Maltese Islands. These originate from different sites, namely from Tarxien (1961, four dates), Skorba (1966, eight dates), Mgarr (1961, one date), from Qala ilPellegrin (1977, one date), from the Xaghra Circle in Gozo (1987-1995, 14 dates), all published by Trump (1995-96: 176-7) and, in 1999, one each from the Hypogeum, the Burmghez Cave and one from a cave near Mnajdra Temples (Mifsud, 1999: 165).8 The carbon dates from Tarxien, Mgarr and Skorba had been determined by the British Museum in the 1960s and provided the first set of absolute dates for the cultural sequence of the Maltese Islands. In the 1960s, the radiocarbon years were not yet calibrated and the results were taken to be equivalent to calendar years. This resulted in a chronology with at times considerably too young dates, particularly for the early phases of prehistory (see Trump, 1966: 48). Once the first correction tables derived from tree-ring studies for the radiocarbon years became available in the early 1970s, the dates of Malta’s prehistoric phases were pushed backwards. Trump’s table published in Harrison Lewis’ book on the Maltese antiquities (1977: 8) firmly dated the arrival of the first settlers to 5000 cal. BC (see also Trump, 1990: 20), a date Trump derived from averaging the 1σ calibration of one radiocarbon date from an Ghar Dalam context in Skorba (ibid., 2002: 23). The number of radiocarbon dates from secure contexts used for the fixing of the chronological dateline has not changed since the 1960s, but the more recent literature shows some variations (see Table 3.1.).

However, radiocarbon date measurements of samples of a known age were found to disagree with their corresponding calendar dates, as radiocarbon years are not equal to calendar years, because of temporal fluctuations in C14 concentrations in the atmosphere. These discrepancies needed to be corrected by calibrating radiocarbon dates to other historically dated material. Measuring radiocarbon dates of sequential dendrochronologically aged trees (mainly bristlecone pines in the US and German and Irish oaks) over the past 10 years has produced a calendrical/radiocarbon calibration curve, which now extends back over 25,000 years (Bronk Ramsey, 2005). This enables the calibration of radiocarbon dates to solar or calendar years. However, the radiocarbon timescale fluctuates up and down and thus the same radiocarbon age can be equivalent to one, two or more calendar years. Furthermore, there is always a measurement error on any radiocarbon date, and it is the overall probability distribution that must in reality be calibrated, not just one age. The calibration process thus transforms the original Gaussian radiocarbon measurement via another set of probability distributions into a calendar age probability distribution, and a computer performs the complex statistics involved, using specialised software (see Reimer et al., 2004).

There are several limitations with regards to the Maltese carbon dates. For a start, many of the standard deviations are very high, resulting in very wide age brackets already when using 1 standard deviation (i.e. with 68.2% probability). Applying 2σ standard deviations, as is current best practice, the age range for the two Ghar Dalam dates becomes 5500-4100 cal. BC (see Table 3.2). Furthermore, the 1 standard deviation leads to several overlaps, which become even more apparent with the 2 standard deviations (see Figure 3.1 and Table 3.1.). Although these overlaps in dates may seem to suggest overlaps and perhaps even co-existences of some cultural phases, this would, in practice, require more investigative studies into these very interesting possibilities, particularly with regards to the end of the Temple Period and the subsequent Tarxien Cemetery Phase.

While the conventional radiocarbon age never changes, calibrated data change with successive calibration curves that further refine the resolution of the fluctuations within the curve. The calibration result depends entirely on the shape of the calibration curve for the relevant period. The original radiocarbon measurement thus can be transformed into a very narrow calendar age distribution, when the calibration curve is steeply sloping, but where the calibration curve is rather flat, the opposite occurs. Furthermore, significant fluctuations in the calibration curve may lead to multiple possible calendar age ranges (see below, e.g. Sample 90). As the 1σ calibration range with 68.2% probability presupposes more accuracy in the data than can be justified, it is current best practise to use the 2σ calibration range, which has a 95.4% probability (Beta Analytic, 2005; Hunt, personal communication. 2006).

The problem with dating wood charcoal (i.e. all BM samples except BM-141, BM-711 and BM-710) is that the date derived may be too early by an unknown number of years (e.g. by dating recycled beams). Also, samples may be mistakenly attributed to later phases. This happens when deposits were disturbed (Trump, 2002: 55) 8 A carbon date from a human bone found in a cave near Mnajdra proved to be of Byzantine age (OxA-8166, 1325 +/- 50 BP converted with a 2 σ standard deviation to 610-810 AD; ORAU Datelist 28). It is, therefore, not considered here.

34

Period

Phase

Neolithic Period

Ghar Dalam Grey Skorba Red Skorba

Years BC (Bonanno, 2000)

Years BC (Trump, 2002)

Years BC (Pace, 2004)

5000-4500 4500-4400 4400-4100 4100-3800 3800-3600 36003300/3000 3300/30002500

5000-4500 4500-4400 4400-4100

5200-4500 4500-4400 4400-4100

4100-3700 3800-3600 3600-3200 3300-3000 3150-2500

4100-3800 3800-3600 3600-3000 3300-3000 3000-2500

2400-1500 1500-700 900-700

2500-1500 1500-? 900-700

Temple Period

Zebbug Mgarr Ggantija Saflieni Tarxien

Bronze Age

Tarxien Cemetery Borg in-Nadur Bahrija

2500-1500 1500-700 900-700

Phoenician/ Punic Period

Phoenician Punic

700-550 550-218

Roman Period

Republican Imperial Byzantine

218 – 30 30 – 535 AD 535 - 870 870 – 1091 1091-1282 1282-1530 1530-1798 1798-1800 1800-1964

Arab Period Post-Arab Period Spanish Rule Knights Period French Period British Period Independent Malta

Since 1964

Carbon date range (Malta) (cal. BC 2σ) 5500-4100 4350-3650 4350-3050 3700-2900 3360-2940

Piano Vento

Carbon date range (Sicily) (cal. BC 2σ) 5300-3500

Lipari Diana

3800-3520

San Cono/ Piano Notaro

3950-3360

Castelluccio

3100-1600

Corresponding type site, Sicily

3500-2470 2900-1420

Table 3.1.: Chronological overview of the cultural sequence in Malta.

as may possibly be the case with BM-142, BM-101, OxA-3572 and OxA-3571 (Trump, 1995-1996: 173-4, and see Table 3.2).

3.3. Chronology, cultural sequence and settlement development in the Maltese Islands There is, as yet, no secure evidence of human settlement in the Maltese Islands prior to the Neolithic Period despite claims of ‘Neanderthal’ teeth from Ghar Dalam and Palaeolithic cave art at Ghar Hassan (see Mifsud & Mifsud, 1997). However, it is possible that prior to the settling of the first colonisers, fishermen from Sicily may have used the islands as a stop-over or even set up seasonal camps during extended fishing trips. In order to find food supply during future trips, they may have deliberately planted some crops and perhaps also introduced some animals, although they may also have introduced some plants and/or animals accidentally. The permanent settling in the Neolithic may, thus, have been the result of a perhaps long series of seasonal visits that extend back to an unknown date. Through the seasonal camps, the landscape may already have experienced some modifications of an unknown extent. During the colonisation by the first permanent inhabitants of the Maltese Islands, sometime between 5500-4100 cal. BC, the possibly already modified landscape was to change its face perhaps even more through the introduction of new animals and plants. From the earliest colonisation phase onwards, there is also archaeological evidence from the island of Gozo that indicates parallel developments with Malta since prehistory. Changes in the cultural sequence in the prehistoric period are usually established through significant changes in the cultural material remains. In Malta, these changes are reflected in the pottery style and through this, Evans (1959) established the first chronological sequence from the pottery in the 1950s.

Overall, the carbon dates from secure contexts available so far cover too few phases satisfactorily. While Trump (1966: 48) formerly gave prominence to the availability of radiocarbon dates in Malta and its potential to export them when Malta’s neighbours had none, the situation has now reversed. Sicily alone has well over 70 radiocarbon dates from different secure contexts (Leighton, 1999: 271-2), and there are many more from South Italy (e.g. Skeates & Whitehouse, 1994). Often, the pottery style of the Maltese settlers may betray their origin. This is particularly so in the Neolithic period, the Zebbug Phase of the Temple Period as well as the Bronze Age phases, as here parallels may be found with nearby Sicily and at times with Calabria in southern Italy (see Bonanno, 2000; Trump, 2002). Here, radiocarbon dates from cultures that share a similar pottery with Malta may be applied to various Maltese phases. As the various timelines of the Maltese chronology (see Table 3.1) are all based on the same set of radiocarbon dates, these radiocarbon dates were re-analysed by the present author using the latest calibration software (OxCal v.3.10) and as a result, a corrected timeline is proposed (Table 3.1). This corrected timeline (with a 2σ standard deviation) will be used hereunder.

35

1σ (68.2% probability) calibrated BC range

2σ (95.4% probability) calibrated BC range

Site- and Lab Code

Archaeological Phase

Conventional carbon date BP

Error (±years)

S-BM378

Ghar Dalam

6140

160

S-BM216

Ghar Dalam

5760

200

-4840

-4360

-5300

-4100

S-BM148

Red Skorba

5175

150

-4230

-3790

-4350

-3650

from -5300

to -4850

from

to

-5500

-4700

X-OxA5038

Zebbug

5330

100

-4320

-4040

-4350

-3960

X-OxA5039

Zebbug

5170

130

-4230

-3790

-4350

-3700

X-OxA3568

Zebbug

5170

65

-4050

-3810

-4230

-3790

S-BM145

Zebbug

5140

150

-4250

-3700

-4350

-3650

S-BM147

Zebbug

5000

150

-3960

-3650

-4250

-3350

X-OxA3567

Zebbug

4860

65

-3710

-3530

-3800

-3510

X-OxA3566

Zebbug

4600

65

-3520

-3120

-3650

-3050

M-BM100

Mgarr

4640

150

-3650

-3100

-3700

-2900

S-BM142

Ggantija

5240

150

-4320

-3820

-4350

-3700

S-BM712

Ggantija

4478

56

-3340

-3090

-3360

-2940

X-OxA3572

Tarxien

5380

70

-4340

-4070

-4350

-4040

S-BM143

Tarxien

4380

150

-3340

-2880

-3500

-2550 -2650

B-OxA8165

Tarxien

4305

65

-3020

-2870

-3150

X-OxA3570

Tarxien

4300

60

-3020

-2870

-3100

-2690

X-OxA3574

Tarxien

4260

60

-3000

-2700

-3030

-2660

X-OxA3569

Tarxien

4250

65

-2930

-2680

-3020

-2620

X-OxA3575

Tarxien

4225

70

-2910

-2670

-3010

-2570

X-OxA3573

Tarxien

4170

65

-2880

-2660

-2900

-2570

H-OxA8197

Tarxien

4130

45

-2870

-2620

-2880

-2570

X-OxA3571

Tarxien

4080

65

-2860

-2490

-2880

-2470

Q-BM808

Temple Period

3912

64

-2480

-2290

-2570

-2200

T-BM101

Tarxien Cemetery

4485

150

-3370

-2930

-3650

-2750

T-BM141

Tarxien Cemetery

3880

150

-2580

-2130

-2900

-1900

X-OxA3750

Tarxien Cemetery

3580

75

-2040

-1770

-2140

-1730

T-BM711

Tarxien Cemetery

3354

76

-1740

-1530

-1880

-1450

T-BM710

Tarxien Cemetery

3286

72

-1670

-1460

-1740

-1420

X-OxA3751

Tarxien Cemetery

1480

70

460

650

420

670

Table 3.2.: Radiocarbon dates from Malta from various sites: Skorba (S-), Xaghra Circle/ Gozo (X-), Mgarr (M-), Burmghez (B-), Hal Saflieni Hypogeum (H-), Qala il-Pellegrin (Q) and Tarxien (T-), analysed at the British Museum (BM) and at the Oxford Accelerator (OxA). Conventional carbon dates after Barker & Mackey (1963: 107-8; 1968: 5-6), Burleigh, Hewson & Meeks (1977: 153-4), Trump (1995-96: 176-7) and Mifsud (1999: 165), 1σ and 2σ calibrations done with OxCal v.3.10 (Bronk Ramsey, 2005). Particularly the 2σ calibrations result in overlaps with several phases as the ranges for the dated phases are considerably wider than conventionally established. S-BM142, X-OxA3572, T-BM101 and XOxA3751 (in italics) are likely to be outliers and, therefore, omitted from the suggested new date ranges. The phase of the date from Qala ilPellegrin is difficult to assess, as Pellegrin ware has been found associated with Zebbug and Ggantija phase ware. The late date suggests a pronounced longevity of the ware type.

This sequence was refined further by Trump’s excavation at Skorba, results of which were published in 1966. Both authors have described the Maltese pottery styles in great detail as the changing pottery styles are very useful to establish the period or phase of the assemblage, and to discover possible origins and connections. Their descriptions will only be briefly touched upon hereunder, as the style plays no vital part in the present study – the few pottery fragments discovered in Marsa Core 1 were unfortunately all undiagnostic. Emphasis will be made here on the cultural developments in the island of Malta in general and in and around the Marsa catchment in particular until the mid 15th century AD, as this is the timeframe covered by the sediments of Marsa Core 1 (see

below). However, as the period between the Knights of St. John up to the present day also includes significant developments, these will be briefly discussed. The developments in Malta will provide the background for the archaeological sites in and around the Marsa catchment, as these may possibly have left some evidence in the sediments of Marsa Core 1. The presence of sites of the various phases is, in the absence of written documentation, based on pottery evidence and stylistic architectural features, but the list is unlikely to be complete as several sites may still await discovery, while possibly no evidence survived from other past sites. Sites that revealed no datable evidence whatsoever are 36

Figure 3.1.: Chronological timescale of the different cultural phases in the Maltese Islands from the Neolithic to Early Modern Malta (Knights Period). Date ranges of most of the Prehistoric phases are given with a 2 σ standard deviation (95.4% probability) when these dates were derived from radiocarbon dates. The arrows denote a stylistic development of the pottery into subsequent phases or ware types. Radiocarbon dates after Burleigh, Hewson & Meeks (1977: 153-154), Trump (1995-96: 176-7) and Mifsud (1999: 165), calibration done with OxCal v.3.10 (Bronk Ramsey, 2005). See also Table 3.2.

pottery discovered within it.9 The cave of Ghar Dalam near Birzebbugia in the southeastern part of the island (see Figure 3.2a.) yielded the greatest number of pottery sherds from the earliest settlers and thus became the type site for this first phase.10 As the pottery found here is

excluded from the discussion hereunder. Among these are menhirs (these may be the last surviving remnant of a megalithic temple (e.g. at Il-Hofra in the north of Malta) as well as be associated with dolmens from the Bronze Age), cart ruts (these have been attributed to the Bronze Age (e.g. Trump, 2002) and to the Roman Period (Bonanno, 2005)), and stone circles. 3.3.1 The Neolithic Period 3.3.1.1 Ghar Dalam Phase (5500-4100 2σ cal. BC) The pottery of the first settlers has been found to bear very close similarities to that of Stentinello in Sicily. The Stentinello culture with its impressed ware was widespread and has also been identified in various sites in Calabria in Italy and Sicily (Leighton, 1999: 61). Radiocarbon dates from the early Stentinello site in Acconia in Calabria (Ammermann, 1985; 6960 +/- 100 BP) reveal a 2σ range of 6020-5660 cal. BC, while Piano Vento in Sicily gives a 2σ range of 5300-3500 cal. BC (Leighton, 1999: 271; 5040 +/- 120 bp), and dates from Capo Alfiere in eastern Calabria give a 2σ range of 49604040 cal. BC (see Table 3.3). The 2σ range of the Maltese dates (see Table 3.1) thus fits comfortably into these timeframes.

Figure 3.2a.: Settlements in Malta during the Ghar Dalam phase in the Neolithic Period, presence based on pottery evidence. Dashed line denotes the Marsa catchment. 1=Ghar Dalam cave, 2=Skorba, 3=Ta’ Hagrat. Data from Trump, 1966; Evans, 1971.

Although the karstic nature of the Maltese landscape gives rise to numerous caves, in Malta only one of these has so far been found to have been used for habitation from the earliest phase onwards, as suggested by the

9 In neighbouring Gozo, two cave series (at Ghajn Abdul and a cave located below Sta Verna Temple) may also have provided shelter for the early settlers, as indicated by the Ghar Dalam phase pottery recovered from the nearby Sta Verna Temple (see Figure 1.5., Lewis, 1977: 19; Trump, 2000: 169). 10 However, no great quantity had been discovered here, and the deposit was mixed with sherds of much later type (Trump, 2002: 28).

37

1σ (68.2% probability) calibrated BC range Context

Site

Stentinello

Piano Vento

( ≈ Ghar Dalam)

Error (± years)

from

to

from

to

A-4474

6130

90

-5210

-4960

-5300

-4840

S-4473

5040

120

-3960

-3710

-4250

-3500 -4350

Capo Alfiere

?

5650

70

-4550

-4360

-4690

?

5450

60

-4360

-4240

-4450

-4070

?

5410

80

-4350

-4070

-4450

-4040

R-180

5000

200

-4050

-3500

-4350

-3350

R-182

4885

55

-3760

-3630

-3800

-3520

Grotta Cavallo

(≈ Zebbug) Castelluccio

Conventional Age (BP)

(Calabria) Diana (≈ Red Skorba) Lipari San Cono/ Piano Notaro

Lab Code

2σ (95.4% probability) calibrated BC range

La Muculufa

(≈ Tarxien Cemetery)

Monte Grande

Rom430

4925

80

-3800

-3630

-3950

-3530

Rom429

4755

75

-3640

-3380

-3660

-3360

A-3957

4080

180

-2900

-2350

-3100

-2000

A-6547

3990

60

-2620

-2350

-2850

-2250

A-3953

3810

120

-2460

-2060

-2600

-1900

A-5283

3790

60

-2340

-2060

-2460

-2030

A-3967

3730

90

-2290

-1980

-2500

-1900

A-5284

3680

100

-2210

-1920

-2450

-1750

A-3959

3640

80

-2140

-1910

-2300

-1750

A-3964

3610

120

-2140

-1770

-2350

-1600

A-3956

3600

100

-2140

-1770

-2300

-1650

A-5724

3700

65

-2200

-1980

-2290

-1910

A-5722

3495

45

-1890

-1750

-1940

-1690

Table 3.3.: Radiocarbon dates from Sicily (after Leighton, 1999: 271-2) and Calabria (after Morter, 1992) from sites with pottery styles that find parallels in Malta during various prehistoric phases. Calibration of the conventional C14 dates with 1σ and 2σ ranges done with OxCal v.3.10 (Bronk Ramsey, 2005). The 2σ calibrated dates from Sicily agree well with the Maltese dates of the corresponding phase.

almost identical with that made and used by the Stentinello culture, one of the earliest farming communities in nearby Sicily, this suggests that Malta was colonised by people from Sicily (Evans, 1959: 45).

they represent an interment (Trump, 2005: personal communication). The three sites in Malta that revealed Ghar Dalam phase material indicate that there was a viable population (Trump, 2002: 29), the size of which was possibly very low (see below).

This earliest phase of the Neolithic period is possibly of considerable importance with regards to the impact these first settlers may have had on the environment, although, so far, only three sites datable to this period have been discovered in Malta (see Figure 3.2.). Apart from at the Ghar Dalam cave, a few pottery sherds were also discovered at Ta’ Hagrat at Mgarr, but this site revealed nothing more from this phase (Evans, 1971: 33). Excavations at Skorba in the 1960s yielded very important evidence as the remains discovered here went beyond the assertion of a mere presence. Remains of a wall from the Ghar Dalam phase hut together with pottery, but more importantly, plant and animal remains indicated that the first settlers introduced the domestic animals cattle, sheep/goat and pigs (Gandert, 1966).

3.3.1.2. Grey Skorba Phase (traditionally dated 45004400 BC) This phase finds no apparent parallels in Sicily and Calabria, and there is no radiocarbon date from a Grey Skorba context at Skorba in Malta. Thus, the date and length of this phase was derived by Trump from his study of the pottery style evolution at Skorba (1966: 28, 50). Grey Skorba pottery is characterised by its grey polished surface and is mostly undecorated. In its early phase, pottery shapes show evolvement out of the Ghar Dalam phase ware, as, for example, handles become higher and more deeply saddled, especially on the ladles (Trump, 2002: 47). Later in the phase, new forms evolve, as for example the trumpet lugs and the presence of some scratched designs (Trump, 1966: 27-28).

From Skorba come also the only human remains from the Ghar Dalam phase. Fragmentary as they are, the jaw and skull bones belonged to at least two children aged about 7 years and about 4½ years, and possibly to a young male (Mangion, 1966). However, their find-spot (from the floor of the Ghar Dalam hut and among stony debris) and fragmentary nature leaves considerable doubt whether

Apart from the excavation site at Skorba, pottery belonging to this phase has also been discovered at the 38

Red slip pottery is also found on Lipari and it may have been trade contacts between the two islands that could have inspired the new look of the local ware, as Lipari obsidian has also been found in the Red Skorba deposits at Skorba (Trump, 1966: 50). There are a number of differences in detail from this Diana ware in shape and decoration (Trump, 2002: 48), but similarities exist with the horn-shaped forked handles of ladles and the characteristic trumpet lugs (Bonanno, 2000: 7). Thus, it is very possible that the red slip of the Diana ware and possibly the above mentioned shapes might have been adapted by the local pottery manufacture.

Figure 3.2b.: Settlements in Malta during the Grey Skorba phase in the Neolithic Period, presence based on pottery evidence. Dashed line denotes the Marsa catchment. 1=Ghar Dalam cave, 2=Skorba, 3=Ta’ Hagrat. Data from Trump, 1966; Evans, 1971.

Ghar Dalam cave and one single sherd was discovered at Mgarr (Trump, 1966: 17; Evans, 1971: 20; see Figure 3.2b.), while in Gozo it has been discovered in a mixed deposit together with imported pottery from Trefontane in Sicily and early Zebbug wares (see below) at Santa Verna (Trump, 1966: 45). The yield at Ghar Dalam cave was very low: only three grey Skorba sherds had been found here, compared to 82 sherds belonging to the previous Ghar Dalam phase11 (Evans, 1971: 20). Considering the relative scarcity of material attributed to this transitional phase and its apparent close relationship to the previous and subsequent phase, a cultural continuity appears implied, but its traditional date may need revision (see below).

Figure 3.2c.: Settlements in Malta during the Red Skorba phase in the Neolithic Period, presence based on pottery evidence. Dashed line denotes the Marsa catchment. 2=Skorba. Data from Trump, 1966; Evans, 1971.

Material remains other than pottery have also been discovered at Skorba, where two huts, one slightly larger than the other, appear to have been used as shrines. The huts consisted of a cobbled surface and remains of a stonewall that could not have been much higher than around 70cm on average, onto which another wall was built with mudbrick (Trump, 1966: 12-3). Fragments of several female figurines were found in one of the huts, which could indicate the veneration of a goddess of fertility or mother goddess (ibid.).

3.3.1.3 Red Skorba Phase (4350-3650 2σ cal. BC) The following Red Skorba phase, presence of which was found only at Skorba, evolved out of the previous Grey Skorba phase and shows a cultural continuity (Trump, 2002: 30). However, the sudden appearance of red slip and trumpet lugs seems to betray cultural influence and parallels with the contemporary Diana culture in Sicily and Lipari (Bonanno, 2000: 7). The Maltese radiocarbon date for this phase shows a 2σ range between 4350-3650 cal. BC (see Table 3.2), while Diana on Lipari has been dated at 5000 +/- 200 bp and 4885 +/- 55 bp (Leighton, 1999: 271), resulting in a 2σ range of 3800-3520 cal. BC and 4350-3350 cal. BC respectively (Table 3.3). Due to the very large standard deviation of one of the Lipari dates, there is concordance with the Maltese date. Nonetheless, according to the 2σ standard deviation given in the Maltese radiocarbon date, the 4400 BC given in the literature may possibly be too early a date for Red Skorba in Malta.

The close proximity of two of the Neolithic sites is similar to the proximity of Stentinello and Diana Phase sites in Acconia in south Italy, which may perhaps also be an indication of close cultural relations (Ammermann, 1985: 95). None of the three Neolithic sites discovered in Malta so far is located within or close to the Marsa catchment (see Figure 3.2). 3.3.2.

The Temple Period

3.3.2.1. Zebbug Phase (4350-3050 2σ cal. BC) One of the most important results of Trump’s excavations at Skorba was the stratified sequence, which showed a clear break in the development between the Red Skorba phase and the subsequent Zebbug Phase (Trump, 1966:

11 The sherds recovered from Ghar Dalam came from a mixed deposit and had been stored at the Museums Department, where Evans (1971: 20) had sorted and quantified them.

39

20). The Zebbug phase is characterised by the arrival of a new group of settlers from Sicily that spread throughout Malta (see Figure 3.3a.) and Gozo. Similarities in the pottery indicate that these people may be related to the San Cono – Piano Notaro culture (Bonanno, 2000: 11). This pottery style is widely distributed, as excavations at Serra del Palco near Agrigento in Sicily revealed. At this site, votive deposits containing small jars with San Cono – Piano Notaro pottery were found packed with ochre. The red variety of the natural pigment appears to have been valued as a magical substance from early times (Holloway, 1999: 17). In Malta, this red ochre was used for painted pottery decoration (Trump, 2002: 50), but more importantly it formed a substantial component of burials as indicated by excavations at Ta’ Trapna near Zebbug, which became the type site for this phase (MAR, 1947-48: 1). More recently, excavations at the Xaghra Circle in Gozo further revealed Zebbug phase burials covered in quantities of ochre (Malone et al., 2005: 79). As ochre does not occur naturally in Malta, a recent study suggests that Sicilian ochre was shipped to Malta in vessels of San Cono – Piano Notaro and Grotta Zubbia ware, which contributed to the new fashion of pottery in Malta (Holloway, ibid.).

chamber tombs as well as in natural caves (e.g. at the Hal Saflieni Hypogeum, which was a natural cave that was extended and adapted in later stages). The position of the inhumed bodies varied greatly, but there is no evidence as yet for cremation of bodies before the Bronze Age (Trump, 2002: 117 and see below). Also, the pottery remains that were found attest to the rapid spread of this culture to a number of new sites in Malta (Figure 3.3a.) and Gozo. The context of these new sites is unclear – they may represent settlement sites, over which, in the subsequent phases megalithic temple structures were erected, but these settlement sites may also have contained shrines, as was probably the case in the Red Skorba deposit at Skorba (see above). From the Zebbug phase at Skorba, the remains of several huts were found, constructed as in previous phases from stone, mudbrick, clay and daub. As some of these contained hearth hollows (Trump, 1966: 14), a domestic context seems probable. Apart from Skorba, also the Ghar Dalam cave in the south of the island and Ta’ Hagrat, which were also occupied previously in the Neolithic (see above), were re-used by the new settlers. It is unclear whether this implies continuity or whether there are gaps in the occupation as could, for example, be suggested by the absence of the Red Skorba phase remains at Ghar Dalam cave and at Ta’ Hagrat. Of all the pottery sherds that have been quantified by Evans, the comparably highest amount of Zebbug phase pottery sherds have been found at Ta’ Hagrat (Evans, 1971: 33 and Figure 3.3a.).12 This is the first phase where archaeological evidence attests the presence of inhabitants within the Marsa catchment, albeit in the form of tombs at Zebbug and at Buqana near Attard, while three other sites that revealed pottery from this phase are located on the fringe of the catchment (see Figure 3.3a.). 3.3.2.2. Mgarr Phase (3700-2900 2σ cal. BC) This phase is generally considered to be a transitional phase as is suggested by the pottery decorations, which are more regular than Zebbug Phase ware, but coarser than the following ware of the Ggantija phase. Mgarr Phase pottery has been discovered in comparatively small quantities in mixed deposits in nearly all the sites that were occupied during the Zebbug Phase in Malta13 and in three new sites, among which are the Xemxija tombs (see Figure 3.3b.). These communal tombs with shaft and chamber also contained burials of the subsequent Ggantija phase. Only the discovery of a pure level during the Skorba excavations made it seem to be a proper phase (Trump, 1966: 38).

Figure 3.3a.: Settlement development in Malta during the Zebbug phase of the Temple Period, presence based on pottery evidence. Sites marked in black denote burial places, in grey sites that were used for habitation and/or ritual activities, outlined circles are possible sites, the dashed line denotes the Marsa catchment. 1=Ghar Dalam cave, 2=Skorba, 3=Ta‘ Hagrat, 4=Hal Saflieni hypogeum, 5=Kordin III, 6=Mnajdra, 7=Tarxien temples, 8=Xrobb l-Ghagin, 9=Buqana rock tomb, 10=Zebbug tombs, 11=Qaliet Marku, 12=Tal Qadi,. Data From Trump, 1966; Evans, 1971, Bonanno et al., 2000.

For this period, there are seven radiocarbon dates from Malta and Gozo (see Table 3.2), which place this phase on a 2σ range between 4350-3050 cal. BC. Although here, OxA-3566 appears anomalously low, the other end of its 2σ range would bring it into line (see also Figure 3.1). Radiocarbon dates from Grotta Cavallo in Sicily give a 2σ date range between 3950-3360 cal. BC (Leighton, 1999: 271-2 and Table 3.3).

So far, this phase has no parallels outside Malta. One radiocarbon date suggests a 2σ range between 3700-2900 cal. BC. The comparatively small quantity of the pottery found may be indicative of the short duration of this phase, especially when considering its wide distribution 12

Meanwhile, however, also the Zebbug phase tomb at the Xaghra Circle in Gozo revealed abundant pottery material, mostly consisting of whole pots (Bonanno, 1999). 13 Only at Tal Qadi has no pottery from the Mgarr Phase been discovered.

The Zebbug phase provides the earliest archaeological evidence of communal burial in rock-cut shaft and 40

stone constructions may also have been a more durable version of buildings that were not preserved (Bonanno, 2000: 14). The largest temple built during this phase, Ggantija, stands in Gozo and became the type site that gave this phase the name (Figure 1.5.). Several other temples appear to have been constructed during this phase in Malta, but on a much smaller scale, namely the east temple at Tarxien, Kordin III, parts at Ta’ Hagrat, the small eastern temple at Mnajdra and a trefoil temple at Skorba (see Trump, 2002). Pottery belonging to this phase has also been discovered at Debdieba, Hal Ginwi and Xrobb l-Ghagin, but it is difficult to associate the sherds with the megalithic remains as the latter are today destroyed (Evans, 1971: 22-24, 26, 27). At Ghajn Zejtuna, the trefoil shaped remains would suggest a Ggantija phase date, but no pottery that could confirm this has been found (ibid., 29). Similarly, at Hagar Qim the shape would suggest Ggantija phase, but here, the pottery recovered would point to the later Tarxien phase (Trump, 2002: 145).

Figure 3.3b.: Settlement development in Malta during the Mgarr phase of the Temple Period, presence based on pottery evidence. Sites marked in black denote burial places, in grey sites that were used for habitation and/or ritual activities, outlined circles are possible sites, the dashed line denotes the Marsa catchment. 1=Ghar Dalam cave, 2=Skorba, 3=Ta‘ Hagrat, 4=Hal Saflieni hypogeum, 5=Kordin III, 6=Mnajdra, 7=Tarxien temples, 8=Xrobb l-Ghagin, 9=Buqana rock tomb, 10=Zebbug tombs, 11=Qaliet Marku, 12=Tal Qadi,. 13=Xemxija tomb, 14=Ghar in-Nghag cave, 15=Hagar Qim, 16=Kordin I + II. Data From Trump, 1966; Evans, 1971, Bonanno et al., 2000.

in Malta, but the wide date range is unfortunately of little use in defining it more precisely. 3.3.2.3. Ggantija Phase (3360-2940 2σ cal. BC) Again, there is a strong relationship between the previous phase and this phase, as expressed in the pottery style, which appears to evolve out of the previous phase to form a distinct style of its own and finds no parallels outside Malta (Trump, 1966: 38). And again, there is a continuation in the use of the sites from the previous phase, combined with an increase in new sites, several of which are located in proximity to the Marsa catchment (see Figure 3.3c.).

Figure 3.3c.: Settlement development in Malta during the Ggantija phase of the Temple Period, presence based on pottery evidence. Sites marked in black denote burial places, in grey sites that were used for habitation and/or ritual activities, outlined circles are possible sites, the dashed line denotes the Marsa catchment. 1=Ghar Dalam cave, 2=Skorba, 3=Ta‘ Hagrat, 4=Hal Saflieni hypogeum, 5=Kordin III, 6=Mnajdra, 7=Tarxien temples, 8=Xrobb l-Ghagin, 9=Buqana rock tomb, 10=Zebbug tombs, 11=Qaliet Marku, 12=Tal Qadi,. 13=Xemxija tomb, 14=Ghar in-Nghag cave, 15=Hagar Qim, 16=Kordin I + II, 17=Ghajn Tuffieha, 18=Ras ilPellegin, 19=Burmghez cave, 20=Busbisija rock tomb, 21=Nadur rock tomb, 23=Hal Ginwi, 30=Tas-Silg, 32=Ghajn Zejtuna, Data From Trump, 1966; Evans, 1971, Bonanno et al., 2000.

Unfortunately, there is only one radiocarbon date and this gives a 2σ range of 3360-2940 cal. BC. This lies below the current published timelines (see Table 3.1.). Burial rites may have changed slightly, as ochre is no longer sprinkled in lavish quantities over the bodies. The number of new burial sites in the shape of both shaft-andchamber tombs and caves increases, while the Hal Saflieni Hypogeum and the Buqana rock tomb continue to be used (Evans, 1971: 6, 59).

3.3.2.4. Salfieni Phase (traditionally placed around 3300-3000 BC) This is yet another transitional phase, where the relevant pottery style emerges out of the previous (in this case Ggantija) phase and adds new elements that are characteristic of the subsequent (in this case Tarxien) phase (see below). Because of the very close relationship to the previous and next phase, it may be considered as part of these sandwiching phases (e.g. Bonanno, 2000: 5). Evans (1971) assumed the pottery style to be an early form of Tarxien, while Trump (1966: 41) discovered pure

By far the most important tangible development of the Ggantija phase is the construction of the first megalithic structures, the vast majority of which were erected on sites that had been in use during the previous phase. Commonly, and also henceforth referred to as ‘temples’, these structures may have been an above-ground reproduction of the underground burial chambers, but the 41

levels with this pottery style at Skorba and Mgarr (2002: 223) and separated it off as a phase in its own right. Pottery of this type has been noted at only three sites in Malta14 and as no occupation gap is suggested between sites used in the Ggantija phase and in the subsequent Tarxien phase, this may possibly indicate that the Hal Saflieni phase could be a parallel occurrence, co-existing with both phases.

Silg). Several temples were also erected on completely new sites (see Figure 3.3d.). It may be because of a possible longer duration of this phase when compared to the Ggantija phase that the highest amount of pottery sherds belongs to the Tarxien phase, but it could also be a reflection of an increase in population, coupled possibly with an increase in activity. Domestic settlements are again very few. Ghar Dalam cave and two other caves yielded pottery ascribed to the Tarxien phase, while a scatter of Tarxien phase sherds was discovered among Ggantija phase material close to the Roman villa at Ghajn Tuffieha (Evans, 1971: 29). Apart from the Hal Saflieni hypogeum and the Burmghez cave burials, no sites belonging to this phase have been found within the Marsa catchment, but several sites are located on its fringe (see Figure 3.3d.).

There is, as yet, no radiocarbon date available for the Saflieni Phase, hence its present time-span is only indicative and may need revision. 3.3.2.5. Tarxien Phase (3150-2450 2σ cal. BC) Three radiocarbon dates from Malta and six from the Xaghra Circle in Gozo limit the 2σ date range to 31502470 cal. BC15 for this last phase of the Temple Period, which is considered as the peak of its development (Bonanno, 2000: 23).

This phase has nothing that is archaeologically comparable in nearby Sicily at that time, and the close similarity in the pottery and building style to the previous Saflieni and Ggantija phases may speak against a foreign influence. The reason or reasons for the end of this phase at around the middle of the 3rd millennium cal. BC are not clear and still the subject of speculation. Whether prolonged drought, over-exploitation of the land (e.g. Bonanno, 2000: 52), or famine, disease and internal struggles (e.g. Trump, 2002: 239-241), or a combination of these factors, the stratigraphic sequence at Skorba and Tarxien and the overall complete change of the material remains suggest that this cultural phase ended abruptly (Evans, 1959: 168; Bonanno, 2000: 52; Trump, 2002: 238).

During this phase, several small temple structures of the Ggantija phase received much larger additions (e.g. at Tarxien temples, Mnajdra temples), or experienced extensions (e.g. Kordin) and alterations (e.g. Skorba, Tas-

3.3.3. The Bronze Age 3.3.3.1. Tarxien Cemetery Phase (2900-1420 2σ cal. BC, Castelluccio culture in Sicily, 3080-1420 2σ cal. BC) Some archaeological evidence suggests that after the Tarxien Phase the island may have been uninhabited for an unknown length of time and that new settlers, again from nearby Sicily, introduced a completely different culture that included, among others, new pottery styles and also new death rituals (e.g. Bonanno, 2000: 52; Trump, 2002: 245). One carbon date, on the other hand, would seem to suggest an overlap of the two cultures or that the new culture did not settle on uninhabited territory and thus could perhaps have ousted the old culture. BM141, from carbonised beans inside a Tarxien Cemetery urn (Trump, 1995-1996: 177), gave a 2σ range between 2900-1900 cal. BC and thus a possible overlap time of around 400 years. Radiocarbon dates from the Sicilian Castelluccio culture, which bears a very strong resemblance to the Tarxien Cemetery Phase, indicate a consistency with this Maltese date particularly from Monte Grande (Leighton, 1999: 271). One date in particular from the Castelluccio culture at La Muculufa, near Agrigento in Sicily, would, at 2σ, put the date to 3100-2000 cal. BC (Table 3.3). The other dates from the Tarxien Cemetery Phase lie between 2140-1420 cal. BC

Figure 3.3d.: Settlement development in Malta during the Ggantija phase of the Temple Period, presence based on pottery evidence. Sites marked in black denote burial places, in grey sites that were used for habitation and/or ritual activities, outlined circles are possible sites, the dashed line denotes the Marsa catchment. 1=Ghar Dalam cave, 2=Skorba, 3=Ta‘ Hagrat, 4=Hal Saflieni hypogeum, 5=Kordin III, 6=Mnajdra, 7=Tarxien temples, 8=Xrobb l-Ghagin, 11=Qaliet Marku, 12=Tal Qadi,. 13=Xemxija tomb, 14=Ghar in-Nghag cave, 15=Hagar Qim, 16=Kordin I + II, 17=Ghajn Tuffieha, 19=Burmghez cave, 22= Debdieba, 23=Hal Ginwi, 24=Il-Qlejgha cave, 25=Borg in-Nadur, 26=Bugibba, 27=Hal Far, 28=it-Tumbata, 29=Kuncizzjoni temple, 30=Tas-Silg, 31=Torri Falka, 33=Tal Lippija, 34=Ta Raddiena, 35=Hal Resqun. Data From Trump, 1966; Evans, 1971, Bonanno et al., 2000. 14

At the Hal Saflieni Hypogeum, which gave this type its name, at the Skorba temple and at Ta’ Hagrat. It may have been present at other sites too, but may possibly have been stylistically attributed to either the Ggantija phase or to the Tarxien phase. 15 The wide date range of the carbon date from Skorba, BM-143, would push the date bracket back to 3500 BC, which is way out of the range of any other carbon date range for this phase (see Table 3.2). Therefore, BM-143, with its 950 year range, is not considered here.

42

(2σ), which fit neatly into the published timelines (e.g. Bonanno, 2000: 5; Trump, 2002: 55). There is no consensus on whether the people associated with the Tarxien Cemetery phase arrived after the island had been deserted for some time (e.g. Bonanno, 2000: 52; Trump, 2002: 245) or whether there are in fact no newcomers at all, but only a radical re-assessment of ideology by the previous people (Stoddart, 1999: 189). As evidence for warfare between the two cultures is lacking, and as nothing appears to survive from the previous phase into the present one, a physical encounter of these two peoples on Malta is generally excluded (e.g. Trump, 2002: 239). However, it should be borne in mind that the calibrated radiocarbon ranges for both phases overlap by more than 400 years (see Table 3.1.). Furthermore, not all warfare need be lethal. The newcomers may have simply exiled the Tarxien phase inhabitants, most likely back to Sicily. The exiling of peoples in their entirety is a common occurrence even in our days, and the exiled people are usually only spared their lives, while the newcomers completely take over their land and possessions but add their culture, which may differ considerably.16 Some people of the Temple Period may also have been subsumed by the Tarxien Cemetery people, but the paucity of material remains belonging to the Tarxien Cemetery Phase would make the number of subsumed people appear very low.

Figure 3.4a.: Settlement development in Malta during the Tarxien Cemetery phase of the Bronze Age Period, presence based on pottery evidence. Sites marked in black denote burial places, in grey sites that were possibly used for habitation and/or ritual activities, outlined black circles are dolmens, the dashed line denotes the Marsa catchment. Menhirs are excluded. 1=Ghar Dalam cave, 2=Skorba, 3=Ta‘ Hagrat, 4=Hal Saflieni hypogeum, 6=Mnajdra, 7=Tarxien temples, 8=Xrobb l-Ghagin, 12=Tal Qadi, 14=Ghar in-Nghag cave, 15=Hagar Qim, 17=Ghajn Tuffieha, 19=Burmghez cave, 22=Debdieba, 23=Hal Ginwi, 25=Borg in-Nadur, 26=Bugibba, 27=Hal Far, 30=TasSilg, 31=Torri Falka, 36=Wied Znuber dolmen, 37=Wied Moqbol dolmen, 38=Misrah Sinjura dolmen, 39=Wied Filep dolmen, 40=Tal Mejtin silos, 42=Il-Bidni dolmen, 43=Ta Hammut dolmen, 44=Birzebbugia dolmen, 45=Ta Firminka dolmen, 46=Ta Hlantun dolmen, 47=Ta Brolli dolmen, 48=Tal Garda dolmen. After Trump, 1966; Evans, 1971; Bonanno et al., 2000; Cilia, (ed.), 2004, Bonanno, 2005.

The Tarxien Cemetery phase saw the introduction of bronze weapons, cremation and a pottery style that betrayed a close relation to the Capo Graziano ware found on Lipari and the pottery of the Castelluccio culture in Sicily (Trump, 2002: 248-9). Nicknamed the ‘Destroyers’ by Evans (1959: 168ff), the new settlers used the Tarxien temples to cremate their dead and also buried them in urns within the walls of the temples. Interestingly, their pottery has been discovered at the vast majority of sites that had been used during the Tarxien phase, including also the Hal Saflieni hypogeum and the Burmghez burial cave (see Figure 3.4.). But whether many of these finds should be qualified as left-overs from casual visits like picnic parties (e.g. Trump, 2002: 251) due to their comparatively small quantity, may perhaps be debatable. Domestic huts belonging to the people of the Tarxien Cemetery phase have been discovered at Borg inNadur and at Skorba (Trump, 2002: 251), while their pottery has also been found at the Ghar Dalam cave, at the Ghar in-Naghag cave (Evans, 1971: 20-1), at Tas-Silg (Bonanno et al., 2000) and at the Xaghra Circle in Gozo (Bonanno, 1999).

dolmens revealed exclusively Tarxien Cemetery pottery and it is thus assumable that all dolmens discovered may belong to this phase (Evans, 1971: 193). The precise function of these dolmens remains unclear, but it is presumed that they have been constructed to house cremation burials, which have, however, not survived (Trump, 2002: 250). Parallels to these constructions may be found at Ognina in southeast Sicily (Leighton, 1999: 137), in the area of Otranto in south Italy, but also in northwest Europe (Bonanno, 2000: 52). No site attributable to this phase has been discovered in the Marsa catchment apart from the Hal Saflieni hypogeum and the Burmghez cave, and only three sites are located just outside the fringe of the catchment (see Figure 3.4a.). 3.3.3.2. Borg in-Nadur phase (ca. 1500-700 BC, Thapsos culture in Sicily 1880-830 2σ cal. BC)

A number of new megalithic constructions in the shape of dolmens are attributed to this phase and they are found at sites that had not been used by previous peoples. The excavation of the Wied Moqbol and Ta’ Hammut

A new wave of immigrants arrived on the Maltese shores and again their pottery style bears close resemblance to a Sicilian culture, this time from Thapsos in southeast Sicily (Leighton, 1999: 192). Evidence from the settlement site at Borg in-Nadur, the type site for this phase, indicates that this culture may have co-existed initially with the Tarxien Cemetery phase. In fact, many of the sites used in the previous phases also contained

16

The exiling of the Jews to Babylon in 587/586 BC may be an early example. After the Second World War, all Germans in Silesia were exiled to the west and their possessions taken over entirely by Polish people. The exiling of the Dalai Lama and many of his people from Tibet, to be replaced by the Chinese culture is a recent example.

43

pottery ascribed to the Borg in-Nadur phase, albeit often in even smaller quantities than in the previous phase.

it could not be linked with any interred body in particular (Evans, 1971: 112). A tomb at Ta’ Vnezja was described as ‘Neolithic’ but a box of pottery labelled ‘Ta’ Vnezja’, which Evans found in the stores of the Museums Department revealed Borg in-Nadur phase pottery. It is unclear whether the pottery from that box originates from that tomb (ibid.: 28). A cave discovered while building an airship station near Qormi revealed Bronze Age material and continued to be used even in Punic times (MAR, 1915-17), but here, unfortunately, more details are lacking and not even the exact location of the cave is known. While the body of a small child has been discovered in the above-mentioned Ghar Mirdum, it is not clear from the report if this represented a burial or an accidental death when the cave collapsed in antiquity due to erosion (see Calleja-Gera, 2001).

Cave settlements remained important and the excavation of a settlement cave at Dingli revealed, among plenty of Borg in-Nadur phase pottery, also an imported pottery sherd from the Castelluccio culture in Sicily (CallejaGera, 2001), which may point to an early use of the cave and again an overlap with the previous phase. New settlement sites are numerous, particularly favoured were easily defended hilltops, and sometimes massive walls are added for protection where the natural defences were considered inadequate (Trump, 2002: 263). A village may also have existed at Tas-Silg, as indicated by sherd scatters in association with mudbrick fragments and stone flooring discovered at the southern slope of Tas-Silg (Fenech, 2001a: 60 and unpublished data), but the remains are too scanty to establish whether the mudbrick construction formed part of an open or a fortified village. Several silo pits, possibly for the storage of water or grain (Bonanno, 2000: 53), have been discovered associated with the settlements (e.g. at Mtarfa and at Borg inNadur), and the presence of silo pits at Tal-Mejtin near Luqa, in which plenty of Borg in-Nadur pottery material had been discovered, could also indicate the vicinity of another settlement.

The exact duration or end of this phase is not known as there are at present no radiocarbon dates from Malta for this phase. However, the silo pits at Mtarfa revealed pottery sherds belonging to the Borg in-Nadur phase as well as to the early Phoenician phase (see below). An overlap of these two cultures may perhaps be implied, although the Phoenician lamp found therein may well be intrusive. Five sites belonging to this phase have been identified within the Marsa catchment, while several more sites are located just outside the catchment (see Figure 3.4b.).

Little is known of the burial practices of the Borg inNadur people. Their pottery has been discovered at one of the Xemxija tombs, in use during the Temple Period, but

3.3.3.3. Bahrija Phase (ca. 900-700 BC) This last phase of the Maltese Bronze Age is not believed to have existed for more than one or two centuries and is believed to be contemporary with the Late Borg in-Nadur

Figure 3.4b.: Settlement development in Malta during the Borg inNadur phase of the Bronze Age Period, presence based on pottery evidence. Sites marked in black denote burial places, in grey sites that were possibly used for habitation and/or ritual activities, dashed line denotes the Marsa catchment. Menhirs are excluded. 1=Ghar Dalam cave, 2=Skorba, 3=Ta‘ Hagrat, 4=Hal Saflieni hypogeum, 5=Kordin III, 6=Mnajdra, 7=Tarxien temples, 8=Xrobb l-Ghagin, 12=Tal Qadi, 13=Xemxija tomb, 14=Ghar in-Nghag cave, 19=Burmghez cave, 22=Debdieba, 23=Hal Ginwi, 25=Borg inNadur, 27=Hal Far, 30=Tas-Silg, 40=Tal Mejtin silos, 49=Bahrija settlement, 50=Ghar Mirdum cave, 51=Il-Qolla silos, 52=Mdina hill, 53=Qallilija, 54=Mtarfa, 55=Qala hill, 56=Ras il-Gebel, 57=Ta Vnezja, 58=Marsa cave, 60=Wardija ta‘ San Gorg, 73=Gebel Ciantar, 74=Ghar il-Friefet. After Trump, 1966; Evans, 1971; Bonanno et al., 2000; Cilia, (ed.), 2004, Bonanno, 2005.

Figure 3.4c.: Settlement development in Malta for the Bahrija phase of the Bronze Age Period, presence based on pottery evidence. Sites marked in grey were possibly used for habitation and/or ritual activities, the dashed line denotes the Marsa catchment. Menhirs are excluded. 1=Ghar Dalam cave, 3=Ta‘ Hagrat, 12=Tal Qadi, 27=Hal Far, 30=Tas-Silg, 49=Bahrija settlement, 53=Qallilija. After Trump, 1966; Evans, 1971; Bonanno et al., 2000; Cilia, (ed.), 2004, Bonanno, 2005

44

phase, but again there are no carbon dates. The new settlers arrived possibly from south Italy and introduced a dark grey to black pottery style with a geometrical pattern, which was infilled with a white paste (Bonanno, 2000: 53). Only one settlement has so far been clearly identified with them, but pottery scatters have been found in several former temple sites, together with other Bronze Age sherds, which may indicate that the Bahrija people did not live secluded from the Borg in-Nadur people (see Evans, 1971: 33; 43; 105; 108 and Bonanno, 2000: 53). Similarly, pottery of the subsequent Phoenician/Punic phase has also been found at these sites (see Figure 3.5., left), which may indicate contact also with these people. The same is valid for the pottery sherds that were found at the Ghar Dalam cave (Evans, 1971: 20).

funerary context in Mtarfa (Bonanno, 2005: 23). This vessel is amazingly similar to a Proto-Corinthian cup that has been found in the necropolis in Motya/Sicily, where it is the earliest datable material that attests the presence of a Phoenician settlement there (Leighton, 1999: 228-9, fig. 121). This striking parallel may reinforce the date postulated by Bonanno for the arrival of the Phoenician settlers at around 700 BC (and not earlier). Also at Mtarfa, in 1939 Ward Perkins discovered a Phoenician oil lamp in a silo pit together with Borg inNadur phase and Bahrija phase pottery (MAR, 1938-39), which may indicate an overlap of the Phoenicians with the Maltese Bronze Age people. A cultural overlap is also indicated by mixed pottery deposits discovered in a tomb at Tas-Sandar and at settlements at Bahrija, Borg inNadur and Tas-Silg (Said-Zammit, 1997: 40). However, the contexts of these deposits are neither sufficiently clear to push the date of the Phoenician settling beyond the late 8th/early 7th century BC, nor to postulate about the nature of a possible co-existence of the different cultures (Bonanno, 2005: 29-35). Having said this, it is interesting to note that Phoenician/Punic pottery has been discovered in many places that have been used by the preceding Bronze Age people, including Ghar Dalam cave and Ghar in-Nghag cave (Evans, 1971: 20, 21 and see Figure 3.5a.). Several rock-cut tombs dated to the early Phoenician period have been discovered in the vicinity of Mdina-Rabat, which may suggest a Phoenician settlement in this area. The variety of burial practices is noteworthy: apart from inhumation and cremation, anthropoid sarcophagi were also used, which possibly reflects the varied backgrounds and origins of the deceased. Unlike in the much earlier Tarxien Cemetery phase, the cremated bones in the cinerary urns were now put into rock-cut tomb chambers or grave-pits (Said-Zammit, 1997: 5).

Little else is known of this cultural phase and so far no burials belonging to this phase have been discovered and material remains other than pottery are also sadly lacking (Trump, 2002: 263). No Bahrija phase site has so far been discovered inside the Marsa catchment, and only one site, Qallilija, is located in the vicinity of the catchment (Figure 3.4c.). 3.3.4. The Historical Period While there may be discrepancies in the dates of the prehistoric periods, there is consensus with regards to Malta’s historical periods, which are divided by the change of political rulers over the islands (see Table 3.1). However, unlike the previous phases and periods, which are divided by a changing style in the assemblages of the cultural material, the change of a political ruler may remain unnoticeable in the cultural material for an amount of time. Nonetheless, these historical dates that mark the changeover of a political ruler are considerably more precise and are shown in Table 3.1.

While the political autonomy in the Phoenician homeland declined, the independence of its colonies in the west grew steadily. Carthage became the most powerful colony until it finally controlled and protected the former Phoenician colonies in the west. Because of these political developments, the period from 500 BC onwards is termed ‘Punic’.

3.3.4.1. The Phoenician/Punic Period (8th century BC – 218 BC) The beginning of the historical period in Malta is determined by the earliest archaeological evidence that testifies the presence of a literate people on Malta – the Phoenicians (Bonanno, 2000: 57). While they may have been visiting the Maltese Islands before that date during their sea travels across the Mediterranean (Sagona et al. (2000: 84) suggest their sporadic presence from as early as 1000 BC), it is presumed that they first settled on the promontory of present day Mdina/Rabat, and adapted the former megalithic temple of Tas-Silg at Marsaxlokk Harbour to honour the Phoenician goddess Astarte by the late 8th century BC (Bonanno, 2005: 23). Securely datable evidence for the Phoenician presence comes from a rockcut tomb at Ghajn Qajjet close to Mdina/Rabat. Here, typically early Phoenician pottery was found together with Greek pottery, namely a proto-Corinthian kotyle, which was datable to the first half of the seventh century BC and an Eastern Greek ‘bird-bowl’, datable to the mid seventh century BC (ibid.). A very fine Proto-Corinthian drinking cup dated ca. 700 BC has been found in a

During this period, there is a marked increase in rock-cut tombs, but compared to the previous period, inhumation is now much more common than cremation (SaidZammit, 1997: 5). Many tombs datable to the Punic period are now also found clustered around the Marsa area, apart from Mdina-Rabat, while other tombs are scattered across the Maltese landscape (see Figure 3.5a.). This may indicate a marked growth of the population, but the appearance of tombs in the Marsa area may point to an increase in harbour activities. A cave that appears to have served for habitation purposes in the Bronze Age was found to have been used subsequently as a burial place, probably in the early Punic period (MAR, 1917-19: XII). The increase in harbour activities may also be indicated by the presence of building remains at the site of the former Technical College in Marsa. This building was originally very large, which may indicate its 45

quite clear (ibid., see also Vidal González, 1996). The majority of Phoenician/Punic sites that are located within the Marsa catchment are burial sites, but several other sites are also located around the fringe of the catchment, Figure 3.5a. Generally, tombs and burial sites appear as if randomly scattered all over the island. The increase in burials particularly in the Grand Harbour area, together with a presumed harbour town called Chersonessos that is mentioned in some editions of Ptolemy’s Geography has been seen in some studies as supporting and even crucial evidence for a town in the Grand Harbour (e.g. Said-Zammit, 1997), but the addition of polis Figure 3.5a.: Settlement development in Malta through the Phoenician/ Punic Period, presence based on pottery to Chersonessos (which and archaeological evidence. The dashed line denotes the Marsa catchment. Only burial places that have already means ‘headland’) is been in use since prehistory are specifically numbered, other burial places are marked out in black squares. 1=Ghar Dalam cave, 7=Tarxien temples, 9=Buqana rock tomb, 14=Ghar in-Nghag cave, 19=Burmghez cave, likely to be erroneous 22=Debdieba, 25=Borg in-Nadur, 27=Hal Far, 30=Tas-Silg, 50=Ghar Mirdum cave, 52=Mdina hill (Melite), (see Bonanno, 2005: 54=Mtarfa, 58=Marsa cave, 59=Ras il-Raheb, 61=Ta Gawhar tower, 63=Zejtun Roman villa, 64=Tal Kaccaturi 202 and ibidem Roman villa, 65=Ghajn Tuffieha Roman villa, 66=Burmarrad Roman villa, 67=Tal Baccari round tower, 68=Tat Ptolemy’s list of Torrijiet, 69=Zurrieq Punic tower, 70=Tad Dawl Roman villa, 71=Wilga tower, 72=Ta Cieda tower. After MAR 1934-35; Evans, 1971; Buhagiar, 1986; Said-Zammit, 1997; Bonanno et al., 2000; Bonanno, 2005. Maltese landmarks), particularly in the absence of any importance. Earliest datable artefacts place it into the 4th archaeological remains in the Grand Harbour area that century BC (MAR, 1955-56: 7). would point to domestic structures (see Figure 3.6.). A large settlement appears to have existed at MdinaRabat, where recent excavations at Mdina revealed that Phoenician remains were replaced and cut into by a massive Punic ashlar wall, which may have been part of a tower. More Punic remains have been discovered in Rabat, which may have been domestic quarters (Cutajar, 2001: 79-85).

3.3.4.2. Roman Period (218 BC – 535 AD) Malta fell under Roman rule at the beginning of the Second Punic War in 218 BC, which brings the Punic period to a close. The archaeological evidence suggests that the material culture took somewhat longer to Romanise, which leads some scholars to describe the Roman Republican phase or parts thereof as Punic Phase IV and V (e.g. Sagona, 2000: 84), Punic Phase V (e.g. Said-Zammit, 1997: 5) or more generally Punico-Roman (see e.g. MAR 1957-58; 1961:6; 1963: 6) or even Romano-Punic for pottery from ca. 50 AD onwards (Sagona, ibid.). How long the original Punic language survived is unclear, but archaeological evidence suggests its survival possibly up to the 5th or 6th century AD: apart from the Punic inscription at the Tac-Caghqi hypogeum in Rabat, Punic letters have also been found in the remaining covering mortar of a baldachin tomb at Agatha

Country houses were possibly scattered around the landscape, although no unequivocally Punic house has so far been discovered. The coutryside is literally dotted with Phoenician and Punic tombs and most excavated Roman country houses were found to be standing on earlier structures, accompanied by Punic pottery (see Figures 3.5a. and 3.5b. and Bonanno, 2005: 91). Several of the so-called ‘Roman towers’ also revealed characteristically Punic ashlar masonry and Punic pottery sherds, but the function of these towers is not 46

discovered at the site of the Roman villas in Zejtun and Birzebbugia indicate their use at least until the 4th century AD (Bonanno, 1992: 27). A fire destroyed the Roman villa at San Pawl Milqi in the late 3rd or early 4th century AD and the site was abandoned afterwards, a fate it shared with the Ta Gawhar tower at around the same time (Bruno, 2004: 166). Excavations in 1881 at Mdina-Rabat, the site of Roman Melite, yielded an example of a fine Roman town house or domus, whose construction is firmly dated to around 125 BC – 50 BC and which remained in use at least until the late 1st century AD (Bonanno, 2005: 160). The Grand Harbour, particularly around Marsa, revealed numerous architectural remains from the Roman period. Many may be connected with intense trading activities in the form of large warehouses, but whether the remains of a Roman bath near Floriana were part of a public or a private bath cannot be ascertained. Numerous tombs and hypogea are prominently scattered around the coastline as if to salute the arriving or departing vessel (see Figure 3.5b. and Figure 3.6.). Although the tombs are often seen to indicate a settlement in this part of the harbour (e.g. Said-Zammit, 1997; Bruno, 2004: 70), the absence of any domestic architectural remains and the former topography of the coastline would make a permanent harbour settlement unlikely.

Figure 3.5b.: Settlement development in Malta through the Roman Period, presence based on pottery and archaeological evidence. The dashed line denotes the Marsa catchment. Only burial places that have already been in use since prehistory are specifically numbered, other burial places are marked out in black squares. 1=Ghar Dalam cave, 7=Tarxien temples, 22=Debdieba, 27=Hal Far, 30=Tas-Silg, 40=Tal Mejtin silos, 50=Ghar Mirdum cave, 52=Mdina hill (Melite), 54=Mtarfa, 59=Ras il-Raheb, 61=Ta Gawhar tower, 63=Zejtun Roman villa, 64=Tal Kaccaturi Roman villa, 65=Ghajn Tuffieha Roman villa, 66=Burmarrad Roman villa, 67=Tal Baccari round tower, 68=Tat Torrijiet, 70=Tad Dawl Roman villa, 71=Wilga tower, 72=Ta Cieda tower. After MAR 1934-35; Evans, 1971; Buhagiar, 1986; Said-Zammit, 1997; Bonanno et al., 2000; Bonanno, 2005.

No. 18 in Rabat (Becker, 1913: 31 and see drawing in Bellanti, 1924: 118). The Roman period spans well over 700 years of Maltese history, but to subdivide this period into Republican and Imperial phases with regards to the settlement pattern and new sites would be fraught with difficulties because at times the MARs are not sufficiently clear and simply refer to a site or the pottery as ‘Roman’ (e.g. MAR 194647: 1683). This long period was, however, not homogenous throughout. Pottery analyses of local and imported amphorae from Tas-Silg and San Pawl Milqi seem to indicate a peak in activity between the late 2nd century BC and early 3rd century AD, followed by a severely marked slump that barely recovers by the end of the Roman period (Bruno, 2004: 112, 132).

With regards to sepulchral architecture, there is at first little change compared to the previous phase and Punic traditions of inhumation, cremation and the use of cinerary urns continues. Then, a rectangular shaft and chamber appears to become the predominant tomb plan, and tombs are much more markedly concentrated around Mdina-Rabat (ibid.: 163). The shaft-and-chamber tomb may gradually have developed into small and larger hypogea that are still scattered around in Rabat and the Maltese countryside. Many tombs and hypogea in the countryside may be associated with the presence of Roman villas in the vicinity, but others around the harbour and not necessarily associated with a settlement may also have been used for non-locals who passed away

There is little change in the general settlement pattern during the early period when compared to the previous phase and a continuation of use of many sites is suggested by pottery finds. Several country houses that appear to have been built in this period are overlying earlier constructions (see above). More tombs and even small hypogea dot the countryside (see Figure 3.5b.) and often seem to be connected with country houses. Coins 47

Figure 3.6.: Topographic map of the Marsa area, showing the location of the boreholes and the various archaeological sites. Altitude contours every 2.5 m. Source: MWPA, MARs 1905-1968, Working Reports of the Government Departments, 1993, Bruno, 2004.

and possibly occupied until the late 9th century (Luttrell, 1975: 22-23). Heavy masonry walls belonging to defence structures were recently discovered during rescue excavations at the Xara Palace Hotel in Mdina and were tentatively dated to the 8th or 9th century AD. These fortifications may reflect insecurity and fears of enemy attacks (Bonanno, 2005: 259).

while on the island. It is well nigh impossible to closely date the tombs and hypogea (see Becker, 1913: 190-193). Archaeological evidence from St. Paul/Agatha hypogeum No. 6 in Rabat may suggest their use from as early as the 1st century AD and its Christian character would have been assumed at a later date. Similarly, it cannot be excluded that Muslims may also have used some of the hypogea for burials (Becker, 1913: 192 and see below).

Figure 3.5b. may show a correlation between villa sites and many burial places, but there is a strong concentration of burial sites on the western side of Mdina hill, where the Roman town Melite was located. Other burial places appear scattered in the southern part of Malta, while there are also quite a number within the Marsa catchment. Archaeological remains of several structures appear for the first time in the vicinity to the coring site in Marsa. They may represent the remains of Roman warehouses and would attest harbour activities. Below the bastions at Floriana remains of a Roman bath were discovered, and several large hypogea were discovered around the present Marsa inlet (Figure 3.5b.). As in the previous periods, no structural remains of a domestic nature were discovered.

In the middle of the 5th century, Malta is likely to have experienced an occupation by the Vandals and in the last quarter of the 5th century an occupation also by the Ostrogoths, and for some time, the ‘barbarians’ were not thought to have caused an obvious break in the Roman life of Malta (e.g. Brown, 1975: 72), but studies by Bruno (2004) imply the opposite (see below). During the 6th century, under the reign of Justinian, Malta was incorporated within the Byzantine Empire together with Sicily. The textual evidence from this period is very scanty and mainly ecclesiastical (see Brown, 1975: 7187), and furthermore, most of the few archaeological remains are of a religious nature. At the Tas-Silg sanctuary, remains of a basilica were discovered and at the rock-cut cemetery of Abbatija tad-Dejr, a shrine was attached. There may have been other Byzantine churches elsewhere, including some rock cut churches associated with early Christian catacombs. Byzantine pottery was scattered around San Pawl Milqi and the villa that had burnt down before (see above) was rebuilt and fortified

3.3.4.3. The Arab Period (870-1091) The inhabitants’ fear of enemy attacks appears to have been justified, as the Arabs conquered the islands from Sicily in the 9th century and took it from the Byzantines. As the Arabs controlled North Africa and also Sicily, 48

widely distributed names, indicative of woodland, crops etc thus probably have their origin at some time after the Arab conquest (Blouet, 1964b: 32). 3.3.4.4. Post-Arab Period (1091-1282) Malta was once more under European domination after Count Roger the Norman took Malta from the Muslims without bloodshed or damage in 1091, and with possibly little other immediate consequences for the island. Unfortunately, no building in Malta may be dated securely to the Norman period, although it is suggested that a crude apsed church and maybe even a monastery may have been built close by the earlier basilica at TasSilg (Luttrell, 1975: 34) but the evidence is so scanty that this is questioned later by Blagg & Luttrell (1990: 123). There is no evidence for a Latin presence at the site of San Pawl Milqi before the 13th century, but the pottery may indicate the continued use of the site in this period by the Muslims (Luttrell, 1975: 34). The 12th century Arab writer Al-Idrisi describes the island as having a good harbour opening to the east, a city and plenty of pastures, flocks, fruit and above all honey during the Arab domination (Amari, 1880: 53). Frederick II brought formidable wealth over Sicily, but it is questionable how Norman Malta really was and thus any parallels to Sicily may be irrelevant. Within a few years after the emperor died in 1250, Frederick II’s empire fell apart and Malta was sacked at least once (Blouet, 1964b: 42). There are no reliable data about site distribution, but it would seem possible that the same sites as in the previous period were occupied. It is also likely that small hamlets may have been scattered around the countryside.

Figure 3.7a.: Settlement development in the Arab Period. The dashed line marks out the Marsa catchment. 30 = Tas-Silg, 52 = Mdina hill, 66 = San Pawl Milqi, 75 = St. Angelo, 76 = Timber Wharf After Luttrell, 1975; Wettinger, 1975; Buhagiar, 1986; MAR 1993.

Malta had only limited strategic value to them, but within this security, the island became quietly prosperous (Blouet, 1964b: 30). Documentary evidence is very sparse, as is the archaeological evidence. The latter may imply the destruction and rebuilding of the buildings at both Tas-Silg and San Pawl Milqi, where the baptistery at Tas-Silg may have been built over by a mosque (Cagiano de Azevedo, 1975: 88-95). At San Pawl Milqi, the Arabs may have installed themselves within the ruins and built a crude fortification around it (ibid.), but the evidence for this still awaits to be demonstrated in scientific detail (Blagg & Luttrell, 1990: 150). The former town of Melite was probably now reduced in size and refortified, as Muslim tombs at the site of the above-mentioned Roman town house were discovered and these would have been located outside the town walls. The area now occupied by Fort St. Angelo in the Grand Harbour was probably also fortified, as the Arabs probably exploited the nearby creek for commercial shipping (Blouet, 1964b: 31). It is also possible that the Muslim population may have continued to use some of the Late Roman/Byzantine hypogea outside the former town of Melite. This suggestion has been put forward by Becker (1913: 193), who saw possible indications for this in the re-use of a Christian altar slab, which, crudely hewn off on two sites, was used upside down as a covering tomb slab in St. Paul’s catacombs. Other Muslim tombs were discovered around Marsaxlokk and also in the Grand Harbour area; at the latter site there were also the remains of a building (Cutajar, 2001: 79) However, the overall number of sites that may be attributed to the Arab period is low (see Figure 3.7a.).

3.3.4.5. Spanish Rule (1282-1530) The house of Aragon became overlords of Sicily and Malta, and Malta was moved to the borderlands of a divided and distracted Europe, often to be sold as a fief to the highest bidder, who invariably exploited the island (Luttrell, 1975: 48-51). Malta, which appeared prosperous in the 13th century, now had to deal with invasions, plagues, bad weather, depopulation, overtaxation, corrupt officials and more. There are, however, also indications of wealth based on cotton and piracy and the construction of churches, convents and palaces in Mdina and Rabat (ibid.: 52). The rural community was scattered in around 60 villages and hamlets that dotted the countryside by 1514, while Mdina remained the fortified capital. Many more small churches, including rock-cut churches were dispersed all over the island (Buhagiar, 1975: 164-6). The northwestern part of the island appears strangely bare of villages, despite the fertility of land in this part of the island is very fertile. The abandonment of this area may have been caused by the operations of the Barbary corsairs (Wettinger, 1975: 191). Similarly, the coastal areas remained uninhabited, with the exception of the small port at Birgu, which was in the vicinity of the fortified castle of St. Angelo. The low-lying areas, like Marsa and Burmarrad were also avoided (see Figure 3.7b.). The appearance and

Under the Arab dominion, a change of language took place, which gives a datum for Maltese place names. Place names that describe transitory objects are unlikely to survive such an upheaval in any great numbers. The 49

international and cosmopolitan lifestyle of the order and other Maltese served the needs of the Knights, but large parts of the rural population lived undisturbed in isolation (Bowen-Jones & Beeley, 1961: 112 and Figure 3.7b.). 3.3.4.7. French Period (1798-1800) Under the Knights, the population grew fivefold and by 1798 the majority of the inhabitants of Malta lived in the group of settlements that the Order laid out or developed close to the harbours. Napoleon, on his way to Egypt, unceremoniously ousted the Knights of St John. This happened without widespread destruction and the French contented themselves with plundering the island of its movable riches (Blouet, 1972: 156-7). The French proved to be vastly unpopular and, confined to the capital city Valletta, they were ousted by the British fleet, together with Portuguese and Neapolitan allies in 1800 (Nehring, 1966: 42). Figure 3.7b.: Settlement development from the Spanish to the Knights Period. Source: Wettinger, 1975.

3.3.4.8. British Period (1800-1964)

The British found a shortage of Maltese personnel that were properly trained in finances, be they of public or private nature and thus the local economy was marked by troubles. The situation was not made better by Wellington’s 1838 statement before the House of Lords: ‘We hold it as an important post, as a great military and naval arsenal, and as nothing more.’ The opening of the Suez Canal in 1869 started a new era in Malta and the area around the Grand Harbour was

disappearance of villages and hamlets in the countryside may also have been governed by fiefs and potential pressure from the fief-holders (ibid., 193-4). Figure 3.7b shows the disappearance of villages (from Wettinger, 1975), but unfortunately it is unclear how long the various villages existed before they vanished. Nonetheless, the distribution shows the scattered presence of several settlements within the Marsa catchment. 3.3.4.6. Early Modern Malta – the Knights Period (1530-1798) With the arrival of the Knights of St. John in 1530 the island gained new importance and took up a new role. From an unimportant agricultural outpost of Sicily it was turned into a bastion against the Ottoman Empire and hence experienced a building boom of fortifications around the Grand Harbour after an attack by the Turks in 1551. This attack was followed by another, more serious attack in 1565. A new capital, present-day Valletta, was built on the Sciberras peninsula in the Grand Harbour, and Mdina, the old capital, lost importance. As the Knights of St. John were predominantly a maritime power, the activities were now centred around the Grand Harbour, which is also reflected in the settlement pattern (Blouet, 1972: 61). The society was, however, heterogeneous as the rural population continued to live in subsistence and did not participate in the life style of the Knights. Thus, while the Knights busied themselves with their own political and economic affairs, a small number of noble or wealthy Maltese became associated with the

Figure 3.8a.: Settlement growth and development by 1842. Source: Richardson, 1961.

50

Figure 3.8b.: Settlement growth and development by 1956. Source: Richardson, 1961.

Figure 3.8c.: Settlement growth and development by Source: EEA 2004.

subjected to major changes. New settlement sites, located closer to the harbour sprouted up and existing sites grew (Bowen-Jones & Charlton, 1961: 119). Army barracks were constructed in many parts of the island such as at Tigne, Pembroke and Hal Far, around which many villages grew into towns. With the risk of pirate attacks now practically nil, coastal settlements were also established and grew steadily, while villages in the north also grew. In 1957, 5% of the land was occupied by buildings, connected by 893 road kilometres (Planning Authority, 1990: N18). Urban expansion continued and new towns were founded at Sta Lucia and San Gwann (Boswell, 1994: 135). Clearly, the island was getting crowded (see Figures 3.8a. and b.).

2004.

3.4. Population estimates The impact humans have on the environment may depend in part on the number of people a given environment can sustain. Of course, one destructively minded person may be worse than, say, one hundred responsible citizens, but nonetheless one hundred responsible citizens have relatively more basic needs (food, shelter, water etc.) than one person and as such, may also have a considerable impact on the environment. Estimating population sizes during the various Maltese periods and phases is important for understanding subsistence, socio-economic organisation and population trends. But despite the availability of a variety of methods, based on, for example, the floor area of villages (LeBlanc, 1971), number of households (Lightfoot, 1994), number of rooms (Hill, 1970) or number of pottery sherds (Kohler, 1978), anyone hoping to find reliable population data for the Maltese Islands prior to 1590, when the first census was held (Cassar Pullicino, 1956: 24), will be disappointed, although this has not deterred others from giving population estimates for various periods. First and foremost here was Renfrew (1973: 168-70), who calculated for the Temple Period a maximum calculation of 11,000 people by dividing the islands into six territories with a temple cluster each and in relation with modern arable land. For the Arab Period, the 15th century writer Al-Himyari, quoting an earlier writer, Al-Bakri, describes in his account that Malta was supposedly uninhabited after the Muslim raid in 870 AD for over 180 years, until around 1049 (Brincat, 1991: 11-14). An estimate by Abate Giliberto for King Ferdinand II in 1240 AD puts the population in Malta at 3800 people, which is presumed to have doubled by 1400 (Cassar Pullicino, 1956: 24). Jean Quintin d’Autun, who arrived in Malta shortly before Grandmaster Villier l’Isle Adam and his

3.3.4.9. Independent Malta The British gave Malta its independence in 1964 and it has been a republic since 1974. The Maltese government under Dom Mintoff encouraged the Maltese to build and own their property and subsidised building schemes. It also embarked on major public housing construction in most localities. Since the late 1970s, the sheer quantity of construction, development and redevelopment for hotels and apartments has eliminated many of the obvious rural boundaries between localities (Boswell, 1994: 137). According to the Malta Structure Plan (Planning Authority, 1990: N18), the built-up land thus increased to 16% by 1983, and reached 25% in 2004, while agricultural land steadily declined (MEPA, 2006: 40 and Figure 3.8c). The development of rural areas is also facilitated by increased mobility from cars as remote areas are easily reached. The road network now comprises a staggering 2300 km, for which more than 246 000 vehicles are registered (NSO, 2002: 55). Malta, thus, stands in stark contrast to any other Mediterranean island. 51

entourage in 1530, estimated a population of more than 20,000 (Vella, 1980: 29). But neither for prehistory nor for the historical period up to Early Modern Malta is there sufficiently accurate evidence to arrive at accurate figures and calculate averages. Other studies have shown that any attempt to establish the population density of a given territory by calculating urban and cultivable areas and counting monuments will be too arbitrary to have any meaning (e.g. Guidoboni et al., 1994: 55). This also seems to apply for the Maltese Islands up to the 15th century AD.

phases than later (ibid.), and their relative importance may perhaps also be seen in an Ghar Dalam phase rim decoration of a bowl that is described as an ‘amusing head of a cow’ (Trump, 2002: 57). Otherwise, animal representations appear to have been rare in Neolithic Malta, where the repertoire was dominated by human figures. The analysis of carbonised grains from Skorba (Helbaek, 1966: 53) indicated that the early settlers cultivated the land with barley, wheat (in the majority Emmer Triticum dicoccum), but most probably also Club Wheat (T. compactum) and possibly Einkorn (T. monococcum) as well as lentils (Lens esculenta), and some weeds introduced with the crop plants (Field Madder (Sherardia arvensis) and Caterpillar-plant (Scorpiurus sp.)). Trees were also used as a resource for domestic fires, as indicated by the charcoal remains from the Ghar Dalam phase level at Skorba (Metcalfe, 1966). The charcoal revealed the presence of Judas trees (Cercis siliquastrum), Hawthorn (Crataegus sp.) and Ash (Fraxinus sp.). The diet is likely to have been supplemented by edible wild plants, fish and fowl, and molluscs (marine and land), (Bonanno, 1986b: 20 and see below).

3.5. Agriculture, local resources and economic development from the Neolithic up to 15th century AD The introduction and development of agriculture on the Maltese Islands is presumed to have irreversibly changed the environment through the creation of fields for crops and the excessive grazing after the introduction of sheep and goats (e.g. Schembri, 1993: 36; Grech, 2001), while agricultural activity would have led to soil exhaustion and also erosion since Prehistory (e.g. Bonanno, 2000: 52; Trump, 2002: 240). Throughout the various periods of Malta’s history, the local population needed to sustain itself with locally produced food through agriculture and animal husbandry. The inhabitants are likely to have supplemented their diet with fish, game, molluscs and edible wild plants. As Malta lacks metal and mineral resources, and sometimes also food and other goods, these necessities were obtained by ways of trade or piracy, depending on the exigencies of the times (see below). Unlike piracy, trade is based on buying, selling or exchanging goods and/or services.17

There appears to be little change in the Grey Skorba phase to the above. The subsequent Red Skorba Phase sees the presence of a few spindle whorls, which indicates spinning of fibres and possibly the manufacture of textiles (Trump, 1966: 34). The nature of the fibres is as much a matter of speculation as is the extent of the activity. Thus, for the Neolithic period, the archaeological evidence from Skorba, Ta’ Hagrat and Ghar Dalam appears to indicate the presence of a few small farming communities that lived in subsistence. The agricultural activity would have been according to the size of the population it had to support.

The exploitation of local resources is likely to have had an impact on the environment and the following section will analyse the available archaeological and literary evidence to establish as far as possible a picture of the environment, agricultural activity, economy and trade throughout the various periods since prehistory. As the sediments of Marsa Core 1 only extend to shortly before the arrival of the Knights of St. John in 1530, the analysis will stop there. However, trade and economy since Early Modern Malta has continuously shaped the Marsa plain and coastline. To link the late medieval marsh with the present industrial and leisure zone of the Marsa, its development is briefly traced.

Contact with Lipari and Pantelleria is suggested by the discovery of obsidian at Skorba, while flint, also discovered at Skorba, originated from Sicily. How intense the trade contact was during the earliest phase is a matter of speculation, because flint and obsidian don’t perish and are easily sharpened anew by chipping off blunt parts. On the other hand, perishable goods may also have been imported, but these would not have survived in the archaeological record. Similarly, in the absence of any valuable raw materials occurring naturally on Malta, the Maltese settlers may have traded perishable items against obsidian, flint, ochre and greenstone. Imported sherds (four in all in Malta) from Trefontane in Sicily and from Lipari, discovered in Grey Skorba and Red Skorba levels at Skorba attest to some form of contact with the outside world, as does the discovery of a Grey Skorba sherd in Stentinello in Sicily (Trump, 1966: 45-6), but again, the contact may have been rather casual.

3.5.1. The Neolithic Period A vast part of knowledge about the Neolithic period comes from the excavations at Skorba, where the analysis of animal bones and carbonised grains from Neolithic layers revealed that the first settlers introduced domestic animals: cattle, sheep/goat and pigs (Gandert, 1966) and probably also the first crop seeds for cultivation. Cattle appear to have been relatively more frequent in the early 17

As defined by the Langenscheidt-Longman Contemporary English, (1981).

Dictionary

of

52

3.5.2. The Temple Period

are given importance at the Bugibba temple, where they decorate an altar slab (see Bonanno, 2000: 34). All these animal representations may have a ritual significance, but may also be an indication of the diet, which here would suggest that an array of wild animals might have supplemented the supply of domestic animals.

A wave of immigration from nearby Sicily around the late 5th millennium BC marks the beginning of the Temple Period. Yet, despite the apparent substantial increase in settlers, little appears to have changed economically. The rather primitive domestic architecture from the previous period appears to prevail at least in the first three phases of this period at Skorba (Trump, 1966). Additionally, the animal bone assemblage at Skorba shows little variation, except that cattle appear less frequently than in the Neolithic (Gandert, 1966) and cattle remains were also greatly outnumbered by goat and sheep bone remains at Xemxija and Ras il-Pellegrin (Bonanno, 1986b: 29). A variety of animals were now also represented artistically on stone (in relief or painted), pottery (incised, applied or painted) or as figurines (clay and stone) and pendants and the animal representations become particularly common during the Tarxien Phase of the Temple Period. The representations may be put into three distinct groups: domestic animals, wild animals (bird, fish and reptiles) and zoomorphic representations, i.e. animal representations to which no species has as yet been assigned. Since the latter at present cannot add to our knowledge of the economy, agriculture and environment, they will not be discussed here.

Querns, grinders and rubbers found in great quantity in many sites may indicate that the Temple Period people’s economy was based on agriculture and also relied heavily on wheat cultivation. The presence of querns inside the temples may suggest that cereal grinding took place within the temple precincts, which may imply a socioeconomic role for the temples (Bonanno, 1986b: 25-6), but may also indicate a domestic nature of the megalithic structures. Other cereals may also have been cultivated, but left no record. The small stone altar at Hagar Qim shows a potted (?) plant on its sides, but it is impossible to define the species. The only clear-cut identification of a plant may be made from a pottery vessel found at Tarxien Temples, where the lower part of the carinated bowl shows pronounced imprints of fig leaves (Trump, 2002: 215). Sophistication and technical achievement reach their peaks in the Tarxien phase, and are well expressed in the high quality of pottery, artistic relief decoration in stone and perhaps also in the manufacture of buttons, bonebeads and other personal ornaments. It has thus been suggested that this may indicate a marked division of labour and specialisation into crafts (Bonanno, 1986b: 33). Spindle whorls made of terracotta may attest a continuation of yarn spinning and possibly textile manufacture (Trump, 1966: 34), but the extent is largely unknown. In fact, weaving, and perhaps any leather processing, may have been a household affair (Bonanno, 1986b: 33). The building industry during the Temple Period made a similarly astonishing development. Domestic hut construction changed little since the Neolithic and possibly only involved the effort of one family unit. Rock cut tombs, however, appear for the first time during the Zebbug phase, and the communal burials may have also involved a communal effort for their construction. Similar communal effort would have been necessary to build the first stone temples, which may have had their precursors in mud-brick and wattle-anddaub constructions. By time, some of these early stone constructions from the Ggantija phase were replaced by the more sophisticated megalithic structures of the Tarxien phase (ibid., 36). The evaluation of the effort involved in the construction of these temples has been the subject of several studies (e.g. Renfrew, 1972; McConnell, 1985; Bonanno, 1986b; Clark, 2004), but little is known what, if any, economic development underlies the building of the temples. The numerical superiority of goat and sheep in bone remains as well as in iconographic representations in megalithic art (see above) has led to the suggestion that Malta’s megalithic development may have been the result of a flourishing textile trade with Sicily and the temples thus would represent ‘monuments to the wool trade’ (McConnell, 1985).

Most representations of domestic animals were found at the Tarxien Temples (stone reliefs) and the nearby Hypogeum of Hal-Saflieni (pottery, pendants). The size, positioning and artistic rendering of the stone reliefs at Tarxien may indicate the importance – and presence – of sheep, goat, pig and cattle, with a numerical emphasis on sheep and goats. Representations of domestic animals were otherwise only found in Hagar Qim and Mgarr, where two handles in the form of a ram’s head, looking very similar to each other, decorate a ceramic vessel (Pace, 1996: 73, figs. F1, F2). Depictions of birds on stone, pottery, clay and cuttle-bone are also quite numerous and were found in more temples than any other group of animals. At Ggantija Temples in Gozo, an ibis incised in limestone was found (Zammit Maempel, 2001: 24-5) and an incised pottery sherd shows a group of crested birds in flight, possibly also birds of passage. From Mnajdra Temples, an unstratified accidental find of an aquatic bird (now lost) was made. Other models of aquatic birds were found at Tarxien Temples, but they are too stylised to allow any specific identification. More diagnostic is the bird in relief between a pair of bull’s horns on a pottery sherd also from Tarxien Temples, showing a partridge or a quail (ibid.: 27-8, Plates 01-15). These representations of birds may perhaps indicate that fowl supplemented the diet. Representations of reptiles are much less common, but here the snake stands out. A very large snake relief decorated an orthostat at Ggantija, while the entrance slab of the South Temple at Mnajdra has a very prominent calcite inlay, which looks like a snake. The meaning of the snake is not clear (Trump, 2002: 115). Other reptile representations (out of clay) are interpreted as a tortoise head (Pace, 1996: 10-11) and a possible lizard from Tarxien Temples (Evans, 1959: pl. 71). Fish 53

Trade is based on the exchange of any local production for other materials and commodities that were not or could not be produced locally. Although exchange of goods between local village communities may have occurred, in a limited area like Malta there appears to be no commodity that can be produced by one community but not by another (Bonanno, 1986: 37). On an external level, exotic items may be easily recognised in the archaeological record. Contacts with the San Cono – Piano Notaro culture in Sicily are likely to have been maintained during the Zebbug Phase for ochre, but as ochre was also used in later phases, albeit in smaller quantities (see above), it may have come from different localities in Sicily. Important trade contacts were maintained with Sicily for flint and lava quernstones, and Lipari and Pantelleria for obsidian, while greenstone axes may have been imported from Calabria in Italy (Trump, 1966: 50; Bonanno, 1986b: 38). It is unknown what goods the Maltese settlers may have offered in exchange for ochre, flint and other commodities (Trump, 1966: 50), but as the imported goods occur naturally in great quantities in Sicily and the surrounding isles, their ‘trading price’ need not have been high at all. There is an apparent decrease in the use of obsidian from both Lipari and Pantelleria in the Temple period; 81 pieces were found at Skorba as compared to 231 pieces from the previous Neolithic period at the same site (Trump, ibid.). This may indicate a decrease in trading contacts (e.g. Bonanno, 1986: 39-40) to only occasional contacts (Renfrew, 1972: 170), while Trump (2002: 211-2) argues that trade continued unabated during the Temple Period.

storage. For subsequent periods, it has also been argued that these ‘silos’ may have been re-used as tombs (as argued by Ward-Perkins, see MAR, 1938-39: vii and discussed in Evans, 1971: 107) or in some cases may even have served as vats for textile dyeing (Sagona, 1999: 23-60). Loom weights and spindle whorls discovered at many Borg in-Nadur phase sites point to textile production, which may have supplied more than an internal market. As textiles rarely survive in the archaeological record, the nature of the fibre is again unknown. Trade with Sicily is implied by the pottery remains, which at times appear indistinguishable from the pottery in Thapsos in southeast Sicily (e.g. Leighton, 1999: 152; Bonanno, 2000: 54), while very few sherds of Mycenaean pottery of the IIIB phase have been found in Borg in-Nadur as well as in Thapsos, which may suggest that Mycenaean seafaring traders could have been in contact with the Maltese (Trump, 2002: 293). The Bahrija phase, which overlaps with the Borg in-Nadur phase, did not reveal further evidence with regards to trade or agriculture during the Bronze Age. At Il-Qlejgha in Bahrija, the 40 or so rock-cut bell-shaped silos that were discovered, possibly used for grain storage, may perhaps also belong to the Borg in-Nadur phase (ibid., 290). 3.5.4. The Phoenician/Punic Period The Phoenicians did not expel and replace the local population, but may have infiltrated imperceptibly into the Bronze Age culture of Malta (Bonanno, 2005: 26). At Tas-Silg, environmental evidence indicates a degradation of the environment, possibly due to an abandonment of the site by the late Bronze Age population and a subsequent take-over by the early Phoenician settlers. This change of use and of users of the site is supported by the archaeological evidence from Area C5 at Tas-Silg (Fenech, 2001a: 14ff). The bone assemblage at Tas-Silg datable to the Phoenician/Punic period shows a marked preference for ovicaprids and marine and possibly land molluscs, while pig no longer appears (ibid.: 24, 34, 40; Schembri et al., 2000: 102-109). The absence of pig may be owing to cultural reasons as pig is not known to have been exploited by the Phoenician/Punic people (Vella, 2001: personal communication18). Ovicaprids, on the other hand, also served a ritual purpose as indicated by epigraphic material, the earliest of which is dated to the 6th century BC. The CIS (1883, Vol. I: 123) states that a lamb19 was sacrificed as a substitute for a human sacrifice (mlk’mr) (in Frendo, 1993: 171).

3.5.3. The Bronze Age Period During the earliest phase of this period, the Tarxien Cemetery Phase, little appears to have changed in agriculture compared to the previous period, but the evidence is very scanty. At theTarxien temples, the cremated remains of Tarxien cemetery urn burials contained the charred remains of wheat, horse beans and peas (Trump, 1995-96: 177), all crops that are likely to have been cultivated in Malta during this phase. However, the dearth of settlement sites dated to the Tarxien Cemetery Phase makes an analysis of the economy difficult (Trump, 2002: 255). The bronze axes and daggers discovered at the Tarxien Cemetery may well be the most famous import to the Maltese islands at that time, but these need not necessary imply warfare and may have been objects of prestige, as these copper items, too flimsy for serious combat, were buried together with the cremated human remains in urns (ibid.).

Pollen analysis of a Punic ash pit indicated a not very intensely used landscape probably consisting of a mosaic of grassland, arable farming for cereals and olives, and areas with small trees and scrub (Hunt, 2000: 114). This, however, does not necessarily mean that agriculture and land cultivation was of no particular importance. In fact,

Relatively more information is available from the subsequent Borg in-Nadur phase. Environmental analyses of Borg in-Nadur phase deposits at Tas-Silg seem to indicate a mud-brick hut settlement with agricultural activity that probably included crop rotation (Fenech, 2001a: 56, 60). Bone remains from Tas-Silg indicate the presence of domestic animals (ovicaprids, pig) but small fish, molluscs and fowl supplemented the diet (ibid.: 40). The presence of many ‘silo pits’ on or close to many other settlement sites of this phase may imply grain

18

Dr Nicholas Vella, Department of Classics and Archaeology, University of Malta. 19 mlk’mr has also been read by Cooke & Garbini as mlk’sr, which would change the meaning into ‘the sacrificed of Osiris’ (Frendo, 1993: 171).

54

the most famous work in antiquity on agriculture and animal husbandry was a 28 volume treatise that was written by the Carthaginian Mago, and not by a Roman, as one may suspect vis-à-vis the importance and esteem Romans attached to agriculture (Barcelo, 2004: 43). As the western Phoenicians turned the hinterland of their colonies into productive agricultural land whenever and wherever possible, also Malta’s agriculture is likely to have undergone several productive changes (Bonanno, 2005: 105-7). The possible Punic structures that underlie most of the Roman villas in Malta’s countryside and their large water cisterns may indicate a more optimised exploitation of the land, as also indicated by the numerous tombs found scattered all over the island (see Figure 3.5.). The presence of walnut pollen in the Late Punic ashpit of Tas-Silg (Hunt, 2000: 113) may indicate that the Punic people had introduced this tree to Malta from the Near East, where it is native.20 Perhaps another possible introduction by the Punic people may have been Tetraclinis articulata, the sandarac gum tree, for its wood was valuable in ship repair, but perhaps more importantly, very highly priced for furniture manufacture in nearby Carthage (Barcelo, 2004: 43). This tree occurs today only in very few places in North Africa, near Carthagena in Spain and in Malta21 – all areas that are known to have been colonized by the Punic people, but not connected geographically by a land bridges during sea level recessions, which could have facilitated its natural spread. Furthermore, no Cupressaceae pollen has been detected in the Bronze Age silo deposit at TalMejtin, while other tree pollen, albeit scarce and consisting entirely of pine, was found (Godwin, 1961).22

found and took what was worth taking, but apparently this did not include the island itself. The attack may have severely damaged Malta’s Late Punic agriculture and economy.24 Exotic items found in the early Phoenician tombs in Malta indicate trade contacts with Egypt, Greece and Syria-Palestine (Hölbl, 1989; Bonanno, 2005: 57-58). As the Phoenicians did not replace the local population in one sweep, these exotic objects appear mainly in tombs that were probably occupied by these foreigners. The pottery of the new colonisers differs from the local Borg in-Nadur style, alongside which it occurs in most sites, before the Phoenician/Punic wares eventually replace the Bronze Age ware.25 While generally the range of pottery shapes and ware types of pottery from the Punic period is striking, the many imported wares recovered at Tas-Silg testify that the temple was drawn into an international sphere through trade and visitors to the temple (Sagona et al., 2000: 97). As Malta was on the Phoenician east-west trade route, merchant ships may have used Malta as a port of call, and seafarers may have visited the Tas-Silg sanctuary to make their offerings. Thus, the temple activity is likely to have contributed significantly to the local trade economy. Archaeological evidence suggests that Malta exported very little food merchandise. Pottery from Malta is documented in Sicily during the 6th and 5th century BC, but by the end of the period it was widely distributed up to Ibiza (Bonanno, 1990b: 215, 217). It is possible that Malta had an export trade of other goods, perhaps textiles, during the Punic period, but neither archaeological evidence nor contemporary literary evidence as yet substantiates this hypothesis. The historian Diodorus Siculus (V:12.1-4) in the first century BC maintains that the Phoenicians selected the Maltese Islands because of their excellent harbours and their advantageous position for trade with the west. More importantly, he attributes the great economic process achieved by the inhabitants to the assistance received from the Phoenician merchants (Bonanno, 1990b: 215)

It is possible that during the Phoenician/Punic period in Malta, garum (a fish sauce) may have been produced and a dyeing industry could also have been present, but evidence is scanty (Bonanno, 2005: 108-9). It has been suggested that the round silo-like structures at Birzebbuga may have been used for textile dyeing, possibly with the famous Phoenician purple dye that is made of murex shells (Sagona, 1999), but again, there is not sufficient archaeological evidence to substantiate this (Bonanno, ibid.).

3.5.5. The Roman Period The remains of around 17 Roman country houses in Malta testify to continuous agricultural activity, and the many oil presses found in them may indicate a thriving olive oil industry, while the minimal quantity of imported oil amphorae found so far may suggest that the local production was usually sufficient to satisfy the local demand (Bruno, 2004: 61). Apart from structural remains, epigraphic and iconographic material from the Roman period adds further details. An ear of corn appearing next to a female head on a Late Republican Roman coin struck in Malta is perhaps the earliest reference in historical times to the fertility of the island

Malta is also mentioned briefly by several classical authors. From Naevius (IV, fr 32), we learn that the Roman army, while passing by Malta in 255 BC, ravaged, burnt and laid waste to the whole island.23 They 20

Walnut pollen are without airsacs and thus can only be carried several tens of meters (Pokorny, 1973) 21 Results of pollen analyses from e.g. Spain, Sicily or South Italy do not show pollen identification lower than family level for the cypress family (Cupressaceae) as these are difficult to distinguish (José Carrión, University of Murcia, personal communication, 2006). Hence, despite the presently very limited distribution of Tetraclinis articulata, its possible former presence in Sicily or other parts of the Mediterranean cannot be ruled out. 22 The absence of Cupressaceae pollen may perhaps also be due the processing method applied by the Cambridge Botany School, which could have led to the destruction of this particular pollen type (Schembri & Hunt, forthcoming). 23 Bellum punicaum, fragment 32, IV. „Transit Melitam exercitus Romanus. Insulam integram urit populatur vastat, rem hostium concinnat”.

24 Orosius in the 5th century AD described the same event (4.8.5): “The general consul Atilius ruined the famous Sicilian islands Lipari and Malta” (“Atilius consul Liparam Melitamque insulas Siciliae nobiles pervagatus evertit”). 25 These sites include e.g. Ghar Dalam cave, Ghar in-Nghag cave, Borg in-Nadur, Mdina Hill and Tas-Silg (see Evans, 1971 and Brusasco, 1993 and compare Figures 3.4. and 3.5.)

55

and the importance of the cultivation of wheat (Bonanno, 1976-77: 393). More revealing are the Chrestion and the Ceres Iulia inscriptions from the Roman Period (CIL, X: 7494 and 7501). The Ceres Iulia inscription is dated to AD 14-37 and originates from Gozo (Vella, 1995: 26). It documents the cult of Augustus and his wife in Gozo and Julia is associated with Ceres. In classical mythology, Ceres was goddess of abundance and agriculture, and mistress of Zeus, with whom she had a daughter – Proserpina. Proserpina, also goddess of agriculture, had been snatched by Pluto from the plains of Enna in Sicily and taken to Hades. From there, she was finally liberated but as she had eaten pomegranate seeds in Hades, she had to divide her life: she was to spend four months in the underworld, during which time the land was barren before she could resurface for another eight months and make agricultural produce grow. Both periods represent the agricultural seasons (Souli, 1995: 34-5). Interestingly, the presence of a temple dedicated to Proserpina in Malta is documented by the Chrestion inscription. Dated between 27 BC – 14 AD, it commemorates the restoration of the columns, roofs and walls of the temple of the goddess Proserpina, which were prone to ruin from old age.26 The cult is strongly connected with Sicily, which was one of the most important grain producers for the Roman Empire. In Sicily, mother and daughter were better known as Demeter and Persephone. The veneration of Ceres-Proserpina in the Maltese Islands may perhaps indicate more than just the importance of corn growing in Roman Malta: it seems to appear that while Ceres, the mother, who stands for all-year-round growth and agriculture was venerated in Gozo, the seasonal daughter Proserpina was venerated in Malta.

It is possible that flax retting was practised, wherever the terrain permitted. The suitable places were, however, very few and limited in size, which led to the suggestion that the necessary material for the textile production may have been imported into Malta and woven into textiles here (Bruno, 2004: 63).29 Regardless of the origin of the flax, linen production appears to have been an important industry in Roman Malta, as Diodorus Siculus mentions the presence of skilled artisans, who weaved linen, which was “remarkably sheer and soft” (Bonanno, 1990a: 31). Maltese cloth is also mentioned several times by Diodorus Siculus’ contemporary Marcus Tullius Cicero in his orations against Verres. Cicero describes how Verres reclines like a king on ‘transparent’ cushions filled with roses (II, 5: 27) and refers to the fabric of the cushions as Melitensis – Maltese cloth, synonymous with fine, high quality material, though the nature of the material here remains unclear.30 Verres was well aware of the value of Melitensis, as the town Melita had produced for three years women’s garments for him (II, 4: 103). If flax was retted locally, Melitensis may have been a rare luxury article (and thus worthy of Verres’ greed) due to the natural scarcity of suitable retting areas. However, the high price of the article may also have been due to the additional cost of having to import the linen in the first place, which, if highly skilled artisans were available locally, would have made it economically worthwhile. With Malta’s inclusion into the Roman Empire, trade contacts with the outside world became wider, and Malta could offer port facilities and naval services (Bruno, 2004: 163). That port activity may have grown in importance is evidenced by the structural remains found around Marsa in the Grand Harbour (see above) and at Burmarrad, where remains of what could have been a quay wall had been discovered in 1991 (Bruno, 2004: 70). The remains discovered on Jesuits Hill in Marsa and documented by Barbaro (1794), may possibly be Byzantine as indicated by the pottery and amphorae, which would place the structures in mid/end 7th to 9th century AD. Other remains at Marsa discovered in the 1950s indicate use between the 4th century BC and the 4th century AD (MAR, 1955-56: 7-8) These structural remains are generally interpreted as warehouses, where imported goods could be stored, but may also point to transhipment of goods from one place to another (Bonanno, 2005: 239). A marble inscription that mentions a ‘statio’ had been found on Jesuits Hill (Abela, 1647: 16-17 and CIL X, 7496). This may indicate an official seat for control and perhaps tax purposes during the Roman Period, but as the inscription is lost today, it cannot be dated closely on stylistic grounds (Bruno, 2004: 70-73).

Vine and viticulture may not have been particularly relevant to the local economy, and the conspicuous presence throughout the Roman Period of non-Maltese wine amphorae seems to indicate that the local wine consumption relied on import (Bruno, 2004: 62). The fertility of the land of Malta is mentioned by Ovid (43 BC – 17/18 AD), who says that Malta is fertile next to sterile Pantelleria27 (Fast. 3. 567-8). However, considering Ovid’s poetic style, one may rightfully wonder to what extent the comparison is poetic license or factual.28 Furthermore, Pantelleria is a fertile volcanic island that does not lack natural water resources, although it may perhaps have lacked Malta’s excellent harbours. Hence, Ovid’s portrayed ‘sterility’ seems to be at variance with the picture of the Pantelleria’s landscape and the economic reality that also emerges from the discovery of various remains of Roman villae rusticae with traces of the Roman system of land division (Bruno, 2004: 59).

29

This is common practice, and was, for example, done in Britain in the Middle Ages: wool from British sheep was shipped to Holland, where it was woven into textiles, to be re-imported by Britain as a highly priced cloth (McConnell, 1985). 30 The material could have been woven out of animal fibre (e.g. sheep wool) or plant fibre (e.g. cotton or linen). Since the contemporary Diodorus Siculus mentions specifically linen in a Maltese context, it is likely that Melitensis was also made out of linen.

26

Translation after H.C.R.Vella, 1995: 25. 27 Fertilis est Melite sterili uicina Cosyrae/insula quam Libyci uerberat unda freti. 28 Ovid’s style features sense-contained lines, a primarily dactylic meter and internal rhyme, particularly in the pentameter, where the syllable before the caesura (the mid-line break) frequently rhymes with the last syllable.

56

Amphorae of Tas-Silg and San Pawl Milqi according to chronological phases (mean percentage values). After Bruno, 2004. local ware

imports

mean percentage values

45 40 35 30 25

are unknown and the economic decline appears somewhat strange, especially as, according to Bruno (2004: 166), the rest of the Mediterranean enjoyed a particular commercial prosperity during the 3rd and 4th century AD (ibid.: 166).

20

The pottery remains decline even further during the 5th and 6th 0 century, but coins indicate that this mid 3rdlate 2nd - late late 1st c BC - late 2nd-4th c 5th - mid 6th c 6th - 7th c AD 8th - 9th c AD 10th - 12th c does not necessarily imply mid/late 2nd c 1st c BC late 2nd/early AD AD AD BC 3rd c AD economic inactivity. During this period, the Central Mediterranean is dominated by Vandals, yet despite Figure 3.9.: Amphorae occurrence during various phases of the Roman Period in Malta as an Malta’s strategic location it doesn’t indication of variations in the economic activity, which may also be reflected in the population figures. After Bruno, 2004. appear to have assumed a role in the maritime and commercial Apart from the temple at Tas-Silg with its offerings from operations of the Vandals. foreign patrons,31 the economy of the islands appeared to be founded, as in Hellenistic times, upon agriculture and In Byzantine times, imports increase again, possibly due the consumption of local produce. The relationship to the economic and military role assumed by Sicily and between the Maltese and pirates is not fully understood, its isles in the traffic system of the Mediterranean. but the wintering of pirates in Malta (see Cicero, II, 4: Amphorae from Greece, the Orient and North Africa 103), may also have contributed to the economy, though seem to indicate Malta’s place on a trade route. The not necessarily always in a positive way. Busuttil (1971) imports decrease again in the second phase of Byzantine suggests that the pirates may have protected the temple of Malta (8th-9th century AD), possibly as a result of a reorganisation of maritime traffic. The Arabs increasingly Tas-Silg and the inhabitants during winter, but make their presence felt in the Mediterranean and Malta considering that sea traffic was at a standstill during may have gained strategic importance for the Byzantine winter,32 the inhabitants may have needed protection, perhaps against a fee, from the pirates themselves. In Empire as implied by the presence of amphorae from the Republican times, Malta’s economy – being far away Near East and the eastern Aegean in Malta. The Marsa from the next land and commercial centres – may have warehouses discovered on Jesuits Hill (see above) have benefited from piracy, slave trade and unregulated also been dated to this phase (ibid.: 168-9). These exchange conducted by exiled people, who should have warehouses, constructed on a promontory, may perhaps been there as a punishment, but this illegal activity may indicate that the formerly much more extensive harbour have been curbed by Augustus in the first century AD could no longer be used. This may possibly because due (Bruno, 2004: 172). to the silting up process of what is today known as the Marsa plain. The silting up process would have reduced Bruno’s 2004 analysis of local and imported amphorae, the water depth there beyond navigable conditions, datable to the Roman and later periods, from Tas-Silg and particularly at the height of the northern part of the San Pawl Milqi indicates flourishing economic conditions present race track, where warehouses had been found in Late Republican times and the appearance of a new (MAR 1959-60 and see Figure 3.6.). palette of Maltese products, as evidenced by locally produced amphorae, is documented in this period at 3.5.6. The Arab and Post-Arab Period various sites in the Maltese Islands (see Figure 3.9.). The economic boom of the late Republican era appears to be Pottery evidence from Tas-Silg datable from mid 9th to the 12th century AD may indicate a continuation of the followed by a decrease in economic activity in Imperial use of the site, but the evidence is inconclusive as to Malta, but there is still a certain economic vitality. In whether Malta was depopulated or not as most Arab mid/late 3rd century AD, the villa at San Pawl Milqi is destroyed by a fire and is subsequently abandoned, while remains were subsequently destroyed (Bruno, 2004: 171). the Tas-Silg sanctuary is also less frequented. However, Thus, the economic activity in Arab Malta is, due to the other rural settlements still appear to be in use. The dearth of remains, hard to assess. The 11th century geographer Al-Bakri and the 15th century writer Alreasons behind this drastic decrease in material remains Himyari (who quotes Al-Bakri and other earlier sources), mention that during the early Arab time when Malta was 31 For example a set of huge ivory tusks that had later been stolen by abandoned (see above), shipbuilders visited the islands Verres (Cicero, II, 4. 103-104). 32 because “the wood in it is of the strongest kind”. Pines, According to Flavius Vegetius Renatus (1585), who died 450 AD, the period when the Mediterranean was closed for sailing appears to have junipers and olives are mentioned by Al-Himyari been traditionally from November 11, by which date shipping either (Brincat, 1991: 11-14) but from what kind of tree this returned to its base or sought the closest safe haven, and then resumed very strong wood came is not specified. According to Alth on March 10.This timeframe was generally respected up to the 15 Idrisi, who writes in the times of Count Roger II, the century (Cutajar, 2006). 15 10

5

57

wood from Malta was exported to the citadel of Scicli in Sicily, where there was a centre for wood and timber trade also from other provinces (Amari, 1880-81: 53). Maltese placename evidence that may date back to Arabo-Norman times would suggest that there were a considerable number of groves, where pine, olive, oak and several other trees grew. These stands could be natural remnants, but may also have been deliberately planted in the past, which could indicate a timber wood industry.

is no clear indication of the importance of Malta’s cotton exports (ibid.: 15-16). Linen was also an important crop, but its production depended much more on the amount of rainfall than other crops. Several famines were prevented through the importation of Sicilian wheat, with a crop failure occurring every two to three years. On the other hand, good years sometimes resulted in surplus wheat and barley being exported to Sicily and North Africa (ibid.: 13-14). Dry-farming (grain and pulses) was predominant, but irrigation was also practised, especially for cotton, which was planted in March and harvested in October. More than 200 place names in Malta and Gozo refer to wells of various sizes, while more than 100 place names refer to ‘gardens’ and ‘springs’ (gnien, gonna and ghajn) for horticulture and the growing of vines. Crop rotation was of utmost importance and laws regulated that cotton (for export), fodder, thorns and thistles (for fuel and food) and corn (for local consumption) were to be grown in intervals wherever possible. Gozo, unlike Malta, was customarily cultivated throughout the year without any rest (ibid.: 19). The statement of a certain Michael Xerri in the 15th century, who claimed that in Malta, illiterate persons believed that the year was divided into two parts, winter and summer, further reinforces the seasonal approach to agriculture in Malta (ibid.: 45). An increase in population and of houses by 1449 led to heavily increased land prices and more pressure on the land, where previously untilled areas and common waste lands were converted into cotton fields where possible. In one case in 1463 the town council agreed that even ‘lu ridumi’, the boulder scree at the base of cliff faces, was to be turned into a meadow (ibid.: 1011). The returns received from the exportation of cumin, in importance second to cotton, were used to cover the cost of importation of timber for beams and shipbuilding (ibid.: 17, 22), indicating that the local market could no longer cater for the demand for timber. Cumin was sowed in fields with a light soil and bitter cumin was more popular than sweet cumin, at least in Abela’s times in the 17th century (see below). Vines were mainly cultivated for local consumption, but imports from Sicily harmed the local market. In the late 15th century, viticulture died out almost completely, perhaps because of increased cotton growing and in favour of Sicilian wines. Viticulture needs a reliable supply of water from springs or streams, and fruit trees were often grown among the vineyards, but what kind of fruit is not known (ibid.: 2324). It may not have played an important part at all in the economy, but there was an orange grove at Dejr is-Saf near Girgenti Valley, west of Siggiewi, in 1549, and olive trees were very rare. Animals were also exported, but how important this industry was cannot be assessed. (ibid.: 29, 36). Agriculture heavily relied on ox-power, while mules and donkeys were pack animals, the latter practically owned by every farmer. Horses were relatively scarce and not utilised for agriculture but for military use and for the nobility in Mdina (ibid.: 36-38). Pigs were not very common and often kept in fields. Every peasant family also owned a few sheep and goats, while large flocks were owned by townspeople, the church notaries and others, and were cared for by shepherds. Sheep and goats were the main suppliers of

The Arabs are likely to have introduced or re-emphasised a number of crops, like cotton and citrus fruits, and with the introduction of new agricultural techniques, irrigation was optimised. But apart from the resulting numerous trading links with Sicily, organised piracy probably also contributed to the economy (Blouet, 1964b: 31). Malta was not destroyed when it came under Norman dominion in 1091 and Al-Idrisi describes Malta in the 12th century as prosperous, with a town, abundant pastures, sheep, fruit and honey (Amari, ibid.). He does not mention any textile industry, which may either have been very small or was started shortly afterwards, for the 1164 inventory of the possessions of Gulielmo Scarsaria mentions the presence of Maltese cotton in Genoa (Abulafia, 1975: 106), indicating that cotton was grown and exported from Malta. (Blouet, 1964b: 30-31). In Norman Malta, the economy was closely linked with that of Sicily and it enjoyed prosperity. Particularly under the imperial ambitions of Frederick II, Sicily became fabulously wealthy. Malta is believed to have shared in this wealth, despite a short interval during which Malta was sold to the Genoese Henri Pescatore, who used Malta as a pirate base against Muslims and Greeks until he was expelled in 1223 (ibid.: 34). There is no written reference to Marsa datable to this period, and the only archaeological remain in the Marsa area is a Muslim burial discovered in 1993 near Jesuits Hill (Working Reports of the Government Departments, 1993: 76). 3.5.7. Spanish Rule Under the house of Aragon, Malta was pushed to the borderlands of a divided and distracted Europe and the formerly prosperous island now suffered from plague, drought, depredation, depopulation and poverty. Wettinger’s archival research through church records and notarial deeds dating from 1467 to the 1530s provides possibly the clearest picture so far of the environment and agricultural activity (1981: 1-48). It emerges that the economy of Malta before the Knights depended on agriculture. A pirate fleet, operating from its port, did not have more than six ships and Malta only had a small garrison of less than 50 men and an even smaller one for Gozo (ibid.: 2). Despite the relatively low population of around 10,000 people (and sometimes less), Malta needed to import varying amounts of wheat, barley and pulses from Sicily, which it paid for through the export of cotton and cumin. Cotton was the most important crop, but there 58

dairy products, meat and wool in the local market with grazing grounds strictly monitored (ibid.: 43-44).

military and strategic role of the Order are among the most relevant aspects of their rule. To suit the order’s needs and military purposes, an enormous amount of labour and finances went into the building of fortifications, the building of the city of Valletta and of port facilities. As the Knights had to be of noble descent, the fabulous wealth of the Order was in large parts the revenues derived from external properties and land holdings, although income was also derived from slave trade and piracy. From this, those Maltese who chose to work for the Order also benefited. As a result of economic pressures, agricultural development under the Knights of St. John was marked by a slow modification of crop emphasis and land-use patterns, rather than by the introduction of any new crops.

While agricultural activity took place all over the island, production at Marsa was particularly varied and buzzing. This former part of a larger harbour appears now to have silted up completely and, as a result of the continuous deposition of alluvium, became valuable agricultural land. The largest land holdings were granted out to influential inhabitants as fiefs and Marsa was the largest and most important of all Maltese fiefs. In Wied ilGonna, where in Roman times warehouses graced a harbour, horticulture was now practised. Its vicinity to fresh water made it ideal for the growing of cotton and its relatively large swampy areas at the head of the valley favoured the retting of flax (ibid.: 3, 4). Flax was not an important crop and the amounts exported were much smaller than those of cotton, but this is perhaps due to the scarcity of stagnant water pools that were ideal for retting. One such stagnant pool was at Marsa, in an area still today known as Il-Menqa, meaning water enclosure or pool (ibid.: 29-30). Marsa also had at least two vineyards, which would have had also fruit trees growing between them. The Marsa Hortus, shown on many ancient maps (see above and Figure 2.8.), contributed substantially to the flourishing of this fief. Unsurprisingly, the income thus received from the Marsa fief was nearly six times higher than that of the other fiefs (ibid.: 6, 23-24).

In the 16th century, local deficiencies of food were largely made up for by imports from Sicily, from where Malta enjoyed duty-free grain rights, but the need for massive food imports was a drag on the economy. The small size of the island and the poor quality of much of the agricultural land remained a congenital weakness of local agriculture and the Great Siege of 1565 left the island in a disastrous state after the pillaging by 30,000 to 40,000 Turks during their four month stay (Blouet, 1964b: 5964). General supply problems were worsened by poor harvests and the necessity of feeding the large number of foreign troops brought to the island and the imported labour for the constructions around Valletta (ibid.: 65). The pressure on the land by the increasing population led to pastureland being converted to arable land, while rough grazing grounds (xaghara and moxa) were partially made cultivable. Cotton production was also steadily increased. Vast amounts of soil were used in building, which denuded the fields and made the farmers desperate in accumulating soil from other sources, sometimes scraping out crevices in the rock but also stealing it back from the bastion walls (Blouet, 1972: 147-8). Various grand masters bought lands and added these to the magisterial estates since the 16th century but it was not until the first half of the 17th century that the Order really began to organise the land resources to exercise direct control on the agriculture (Blouet, 1964b: 71).

In the 16th century, shortly before the arrival of the Knights of St. John, cotton, cumin and honey were traded against corn, while water and timber were scarce (Bowen-Jones & Beeley, 1961: 108). Frequent raids by corsairs considerably added hardship to the inhabitants, who described themselves as living “on a rock in the middle of the sea far from help and comfort” (R.M.L., 9, F.108, as quoted in Blouet, 1964b: 45). In 1525, some four hundred Maltese were removed during a single raid and many Maltese who had the means moved to Sicily. From the prosperous appendage that Malta had been by the 13th century, it had slumped to becoming an impoverished and partially abandoned out-post (ibid.: 46).

In 1534, the funeral journey of the first Grand Master of the Knights of St John in Malta, Philippe Villier l’Isle Adam, is described to have followed along a waterway that supposedly lead from Qormi to Marsa before proceeding to Fort St. Angelo for interment (Gambin, 2004). This waterway may be the synonymous with one of the waterways also shown on the map by Quintinus (see Figure 2.8.). A more detailed description of the environment, resources and activity around Marsa is given in the 17th century work Della Descrittione di Malta, by Fra G. Abela. Several things appear to have changed in agriculture since the arrival of the Order. Although Abela describes at length the history of linen, flax was apparently not retted during Abela’s time (ibid.: 125-138). Cotton was still grown but the production was much less than before. Cumin and anis were still grown for export at the same rate, but there was a notable increase in fruit (oranges, citrons, lemons) production,

3.6. Changes and development in the Marsa area since Early Modern Malta Neither agricultural practices, nor the effects of trade, economy, natural or any other events that may have occurred since Early Modern Malta are reflected in the following analyses of sediments from the Marsa Sports Ground. Nonetheless, as the coastline has since experienced serious changes and alterations, which are quite well documented from the 17th century onwards, these will be briefly discussed, to bring the changing history of the Marsa shoreline to the present day. The various cultural aspects resulting from the presence of the Order of St. John in Malta have been the subject of many studies (e.g. Schermerhorn, 1929; Hughes, 1956; Blouet, 1964b; Mallia-Milanes, 1993; etc.), but the 59

some of which were also exported to Sicily. Wine was produced in little quantities and more for the knights than for the local population. Wheat was more important, as was animal husbandry to supply the knights with abundant bread and fresh meat. Gozo produced one third of all produce and was thus predominantly still agricultural, while the Maltese adapted to the different lifestyle and exigencies of the knights.

Nuovo), the inlet on the southern side of Il-Kortin and Marsa was drained and reclaimed. A narrow canal, large enough only for small boats, was retained, into which the excess rainwater was to flow into the sea (ibid.: 404). The innermost and most protected inlets within Marsa, which probably gave it its name in the first place, were thus completely removed. Barbaro (1794: 22) described his outrage concerning the activity at Gezira (today Albert Town): “Quest’ isoletta mutò al presente di condizione per essersi ridotte in secco le acue, che la circondavano”. He also mentions that the new canal went into an older one and that numerous archaeological discoveries were made (see above). After the expulsion of the Knights and the take-over by the French in 1798, an entrenchment at Marsa was used to keep the French within the fortifications (Guillaumier, 1987: 406).

The fief at Marsa was still very important, although it indubitably suffered seriously during the Great Siege in 1565, when the Turks set up their camp there and took advantage of the water supply from the wells until dysentery broke out (Bradford, 1964: 60). Acquired by Grand Master La Cassiere in 1582 for 11,000 scudi, the fief was subsequently leased out to favoured members of the Order (Blouet, 1964b: 71 and 1972: 106). Once the aqueduct brought water directly from the Rabat uplands to Valletta from 1614 onwards, ships no longer needed to proceed to Marsa for their supplies of fresh water (Abela, 1647: 93). According to Abela (1647: 92), an area referred to as Curmi (today Qormi) was slowly developing although the inhabitants suffered from ‘bad air’ and were often pale and deformed. The situation apparently improved once many bakeries sprouted up and big ovens were installed to supply the knights in Valletta with bread. Abela reports that in his time, the air was very pure and Qormi was blessed with many gardens and orchards that extended along an abundant freshwater stream through the valley all the way to Marsa. The ‘bad air’ may possibly be synonymous with ‘mal aria’, the disease associated with stagnant water pools in which the malaria mosquitoes live and breed. However, as the sewage of Valletta and the Three Cities was flowing directly into the Grand Harbour (Cassar et al., 2005: 37), adverse weather conditions may have also contributed to soil and water contamination with faeces in the low lying land at Marsa and Qormi, which may have aggravated the ill health of the area.

3.6.1. The French Period By the end of the 18th century, Malta drew its wealth mainly from the foreign income derived from the Order’s mainland estates, taxation of the Commanderies and from the odd special grant. Exports consisted mainly of cotton, but the imports of staple commodities and a range of superior consumer goods for the Knights dwarfed the significance of the revenue derived from cotton export (Bowen-Jones & Beeley, 1961: 115). The confiscation of the Order’s estates in France from 1792 onwards contributed to the depletion of the external resources, and thus the Order was on the verge of bankruptcy when it was ousted by the French. The impact the French had on the local economy was very negative as the French administration lacked a sound financial basis and was founded on the principle of looting the conquered land (Blouet, 1972: 159). 3.6.2. From the British Period to present day

In the middle of the 17th century, an elaborate plan to build a sea wall across part of the Grand Harbour in Marsa was drawn up in an effort to reclaim all land lying behind it for agriculture. Although the plan was never completed, it resulted in some drainage canals being dug into the marshland at Marsa and part of the reclaimed land was used for cultivation (Blouet, 1972: 147-8). The Marsa fief in the 18th century was still leased out to favoured members of the order, but at ridiculous sums. The annual fee varied from a gun (under Vilhena, 172236), a bunch of flowers (under Pinto, 1741-73) to a horse with a saddle (under de Rohan, 1775-97) (Guillaumier, 1987: 405), perhaps reflecting rather the decadence of the Order than the real value of the fief.

By the time the British expelled the French in 1800, the last important Maltese export market for cotton, Spain, had been lost, but this loss was quickly made up for by an increase in civil port activity and entrepot trade during the Napoleonic Wars. The island became increasingly dependant on external economic activity and the new activities also led to more people working in the service industry rather than with procuring local material resources. The ensuing boom was accompanied by internal inflation and many farmers turned from cotton growing to producing foodstuffs, which fetched increasingly higher prices in Valletta. But the boom was short-lived and with the end of the Napoleonic Wars, Malta took time to adjust to peace and to the growth of an international economy in which Britain led the way (Bowen-Jones & Beeley, 1961: 115-119).

The harbour at Marsa, however, was highly valued and developed. In 1768, Grandmaster Pinto spent vast sums when he ordered the construction of 19 silos and a church at the bottom of Il-Kortin. The area known as Il-Menqa was reclaimed through the dumping of limestone debris. A wide road was cut out of rock in the middle of IlKortin and Marsa Grande (afterwards called Porto

With the opening of the Suez Canal, Malta became an important base for ships and its port facilities needed to be extended. In 1861, the port at Marsa, located at the innermost part of the Grand Harbour, was revamped for new purposes under Governor Le Marchant. The land reclamation of the Knights is likely to have resulted in unwanted flooding of the Marsa plain after heavy 60

rainfalls as the natural drains were curtailed. All plains, pools and marshes were removed in 1865 and storm water was channelled into an artificial canal that led to the sea. Porto Nuovo and Il-Menqa, desiccated through land reclamation under Grandmaster Pinto, were dredged to be used again for maritime activities. This was followed by the building of new quays and the new seaport, Marsa il-Gdida, began to thrive. A breakwater was constructed along Jesuits Hill, to protect the innermost harbour from the north-easterly wind Gregale. In 1875, the foundation stone for Albert Town, located on the former Gezira (see above), was laid to provide the ever-increasing port workers with habitation quarters. On its highest point, the abattoir was built (Guillaumier, 1987: 406-7). Successive interventions on the coastline were aimed to stabilise and establish a concrete coastline and deepen the seabed to allow large ships to enter the innermost parts of the harbour. Since the construction of Malta Shipbuilding between 1976 and 1984, the shoreline at Marsa has remained unchanged till this day (ibid.: 418 and see Figure 1.3.). The former Marsa fief became the property of the British War Department. Its agricultural value either decreased considerably, or horse racing was perhaps valued as more important, for in 1869, a huge race track for the use of military and naval officers, was constructed in the plain, putting an end to any agricultural activity in this area. The construction of a golf course within the boundary of the racetrack in 1888 removed any remaining traces of agriculture. In 1902, the War Department granted the vast area of land to the Malta Sports Club for the recreation of British servicemen for a token one shilling a year for ninety years. More land was added in 1906 and 1908, but the War Office retained the right to use the lands for military purposes (ibid.). Today, the area is occupied by the Marsa Sports Ground, to which meanwhile another race track has been added, as well as tennis courts, a football pitch and several other facilities.

61

the construction of the Malta Shipbuilding Yard in 1976.2 This site is located ca. 600m to the west of the coring point of Marsa Core 1. At little less than one km distance to the northwest, a site investigation also carried out by Harrison & Co in 1997 produced further valuable information with regards to the lithology and stratigraphic sequence at Il-Menqa (Figure 4.1.).3

Chapter IV STRATIGRAPHICAL INVESTIGATIONS 4.1. Introduction A detailed study of Marsa Core 1 forms the basis for the bulk of the knowledge of the stratigraphy of the study area. A correlation with Marsa Core 2 is attempted with simplified logs. Additionally, the test drilling results of three cores done by Harrison & Co in 2000 at the site are listed. Although a detailed description of these test cores is lacking, they can be securely correlated at mean sea level due to the exactly known site level, which since the length of the cores is known, then provides approximate information about the bedrock level. Furthermore, a few points of reference extracted from the Harrison report strengthen the correlation and provide some additional information about the topography and sedimentation.1

4.2. General stratigraphy inferred from borehole surveys Apart from the results of Marsa Core 1, the comparatively less detailed stratigraphy and results obtained from the other boreholes provide important information about the general morphology and spatial disposition of the sedimentary infill in some parts of the Marsa plain and at Il-Menqa. Despite the comparative paucity of boreholes and the lack of rock contour data from the Marsa Sports Ground and the Il-Menqa area, valuable information can be derived from a comparison with the Malta Shipbuilding cores, supported by the Maltacom cores.

Also presented are two selected transects from the log data available from ground investigations done prior to

Figure 4.1.: Location of the different transects in the Marsa/ Grand Harbour Area and of the two cores (Marsa Core 1 and Marsa Core 2) taken during the present study. Source of base map: Malta Environment and Planning Authority (MEPA)

1 The coring was done on 25/04/2000 for the Ministry of Education by Harrison & Co to establish the depth and strength of the bedrock of the site for the purpose of building a sports complex. The coring method employed did not allow recovery of any material. Ground descriptions were limited to short terms like ‘clay’ or ‘water encountered’, but also ‘bad smell’. Although the exact depths are given, these log entries can only be roughly correlated with the present core. Copies of the short drilling report were made available by the Malta Environment and Planning Authority.

2 I am very grateful to Malta Shipbuilding for having made this data available for the study. 3 I am very grateful to Nathaniel Cutajar of the former Museums Department, for making the report available.

62

The borehole data available were sufficient to produce four different transects, which have been plotted to scale and correlated (Figures 4.2., 4.3., 4.4. and 4.5.). The two transverse sections at Malta Shipbuilding run roughly along the longitudinal axis of the former inlet into which the dock of the shipbuilding yard had been constructed. The transverse section at the Maltacom Building at IlMenqa consists of two cores that had been retrieved from under the floor of the building that formerly occupied the site. These two cores run across the longitudinal axis of the inlet. The transverse section at the Marsa Sports Ground, which contains the information of Marsa Core 1, Marsa Core 2 and the Harrison Cores 1, 2 and 3 has a horseshoe configuration, but the general orientation of the transect is roughly perpendicular to the direction of rainwater runoff/freshwater flow into the sea (see Figure 4.1.).

minerals and the production of new minerals by weathering and pedogenesis (Gale & Hoare, 1991: 151), but also under the influence of fire (Jordanova et al., 2001: 1137). Although it can also be used for lithostratigraphic correlation and differentiation within a small region, the original colour of a sediment may have been modified in one particular place, but not in another (Gale & Hoare, 1991: 154-157). Only for Marsa Core 1 and Marsa Core 2 was the colour of the sediments in every sample determined by comparison with the Munsell Soil Color Chart. The Munsell scheme was developed in the United States in 1905 and is the most sophisticated and widely used of several methods of colour quantification as it employs three factors (hue, value and chroma) to make up a solid colour. Their combination forms a numerical notation for each colour. At Malta Shipbuilding and at the Maltacom site at Il-Menqa, the sediment colour was broadly described in terms of ‘black’, ‘grey’ or ‘brown’. Although this is very basic, it still provides important information. There is no colour description for any of the Harrison cores from the Marsa Sports Ground.

As colour is perhaps the most obvious property of sediments, its objective determination is essential for characterisation. Furthermore, the colour of a sediment may also provide indications of the mineralogy as well as vital clues to the origin and environmental history of a material (Gale & Hoare, 1991: 147). The soil colour may also be a useful indicator of humus content (dark), iron content (reddish) and speed of deposition or leaching (pale) in a soil (Evans, 1978: 67) as the original colour may change due to loss, movement and concentration of

The particle size descriptions used for the Malta Shipbuilding cores could be adopted for the description of Marsa Core 1, as sufficient particle size data were here

Figure 4.2.: Graphic representation of the stratigraphy and correlation of Marsa Cores 1 and 2 and the three cores removed from the Marsa Sports Complex by Harrison & Co during geotechnical studies in connection with the construction of the Marsa Sports Complex. The stratigraphy of Marsa Cores 1 and 2 is based on soil colour changes established with the Munsell Soil Color Chart, and for Harrison 1,2 and 3 on the excavation log, using the terminology used in the log. Correlation between all cores was done by the present author. Width of the cores is not to scale.

63

available throughout the core (compare Figure 6.1.).4 Marsa Core 2 generally lacks the scientific determination of the particle size, but a rough description of the texture was made while the samples were being divided for the study.5

accumulations appear to be confined to hollows (see Figure 4.3., Core No. 9 and Core No. 18) or when trapped by rising topography (see Figure 4.4., Core 5, compare Figure 4.6.). This type of deposit may have amassed naturally through soil formation processes, and/or erosive processes. Once the hollow or trap was filled, excess material would then have been washed further downstream by water.

Underlying all cores studied in this survey is Lower Globigerina Limestone. However, in relation to the mean sea level, bedrock was encountered at considerably varying depths, from around 2.6m6 to as far down as 12.1m7 below present sea level. This emphasizes the topographic variations of the underlying rock contours, despite the superficially plain and apparently homogenous topography at the various coring sites (Figure 4.1.). A site investigation of the rock contours at Malta Shipbuilding highlights this even more (Figure 4.6.).8 Comparing the rock contours with the stratigraphy of the boreholes from Malta Shipbuilding, it appears that the topography determines which sediments accumulate, and at what rate. Instances, where dark brown sandy soils were detected as the lowermost deposit directly overlying bedrock, as also found in Marsa Core 1 and Marsa Core 2 at the Marsa Sports Ground, were comparatively rare: only three out of a total of the 29 Malta Shipbuilding boreholes contained this sort of sediment, which is most likely of Quaternary origin (see below). These

The presence of black and black/grey silty clay may be related to the water depth and to quiet waters that inhibit the transportation of oxygen via currents, which leads to anoxic or hypoxic conditions. At Malta Shipbuilding, these dark deposits were only found in four of the 29 cores, all at depths greater than 6m below sea level and in sheltered locations towards the mouth of the inlet that leads into the Grand Harbour basin9 (compare Figures 4.3. and 4.4.). Shallower water deposits, either directly overlying bedrock (e.g. Core No. 17) or overlying black silt (e.g. Core No. 2) or brown sand (e.g. Core No. 9), consist of a grey sediment with varying particle size fractions. This is also a very common deposit within the Marsa Cores 1 and 2. At Malta Shipbuilding, towards the head of the inlet, the grey deposit appears poorly sorted with particle sizes ranging from clay to boulders. This grey material then turns to mainly clay-sized particles towards the mouth of the bay (Cores No. 5 and 6 on Figs. 4.4. and 4.3.) while towards the mouth of the Grand Harbour basin, the particle size increases to silt size with traces of sand. At Malta Shipbuilding, brown silty or sandy clay with limestone boulders overlies all grey deposits and also forms the uppermost deposit of all cores. This layer is very likely to be the result of the land reclamation and restructuring works undertaken by Grandmaster Pinto in 1768 in this part of the harbour, as a result of which the peninsular character of what is today called Albert Town was removed (see above). The top part of the grey silty clay layers that immediately underlie this artificial fill could thus be dated to 1768. The Maltacom transects also show that a

Figure 4.3.: Graphic representation of Transect 1, showing the stratigraphy, bedrock level and correlation of five cores excavated by Malta Shipbuilding during geotechnical studies in connection with the building of the shipbuilding yard in 1976. The stratigraphy is based on the excavation log, using the terminology used in the log. Correlation between the cores was made by the present author. Width of the cores is not to scale. 4

The terms used for the description of particle size fractions is similar to the terms used by Gale/Hoare, 1991: 61-62. Thus, clayey sand for example represents a sample that contains between 50 and 90% sand, and up to 10% clay. 5 Division and texture description was done by Frank Carroll. 6 Maltacom Cores A and B at Il-Menqa. 7 Malta Shipbuilding, Core Nr. 2 in Transect 1. 8 The rock contour investigation had been done prior to the construction of the shipyard for the purpose of finding the most suitable site for the dock in the former inlet. Searches to find similar studies in any other part of the Marsa area proved futile.

9 Apart from the three cores shown in the section that have this kind of deposit, it was also found as the lowermost deposit of Core 20 from Malta Shipbuilding at a depth of 7.93m below sea level. According to the coring location, the whole core should have had a depth of more than 16m below sea level before striking bedrock, however, the record only showed 9.03m of deposits. Due to this discrepancy, this core was omitted from the present study.

64

4.3. Detailed description of the stratigraphy of Marsa Core 1 and Marsa Core 2 from the Marsa Sports Ground The location of Marsa Core 1 and Marsa Core 2 is shown on Figure 4.1. The site for Marsa Core 1 was chosen to be 1m to the south of where a test core had been taken in April 2002. Preliminary investigations from this test core had revealed undisturbed deposits underneath a thin layer of disturbed topsoil, thus the possibility of retrieving an undisturbed sequence was deemed highest here. To get a broader picture of the stratigraphy of the plain, the location of the borehole for Core 2 was chosen at a distance of 43m to the north of Core 1 (see Figure 4.1.).

Figure 4.4.: Graphic representation of Transect 2, showing the stratigraphy, bedrock level and correlation of four cores excavated by Malta Shipbuilding during geotechnical studies in connection with the construction of the shipbuilding yard in 1976. The stratigraphy is based on the excavation log, using the terminology used in the log. Correlation was done by the present author. Width of logs not to scale.

grey sediment is overlain by a similar artificial fill (see Figure 4.5.). Here, this fill is likely to date to 1865, when the British made extensive alterations in the Grand Harbour (see above, Chapter III).10 The cores from the Marsa Sports Ground lack this deposit, but it is present in the section further up, albeit not throughout. It is likely to be directly linked to the immediately overlying artificial clinker layer (see Figures 1.4.1.-3.). Both layers probably acted as a stable bedding with good drainage properties and would thus date back to 1869, when the race course was constructed (see above).

The stratigraphic sequence was established through variations in the moist soil colour, based on the Munsell Soil Color Chart (1994). Colour determination of the moist sediments of Marsa Core 1 and Marsa Core 2 was done by Frank Carroll under natural daylight conditions, while colour determination of the dried sediments of Marsa Core 1 was done by the author in the laboratory under natural daylight that entered from a window, but away from the glare of the sun. The dried colour was often found to vary from the moist colour in chroma, value and/or hue. The extent of the colour change of the sediments from the field state to the dry state may help to establish the degree to which they have undergone postdepositional alteration (see Appendix I). The stratigraphy of Marsa Core 1 was as follows (with the standard Munsell notation in parentheses): 0-20cm 20-30cm 30-50cm 50-100cm

Figure 4.5.: Graphic representation of the stratigraphy and correlation of two cores removed from Il-Menqa at Marsa during geotechnical studies in connection with the construction of the new Maltacom building. Cores were excavated by Harrison & Co and the stratigraphy is based on the excavation log, using the terminology used in the log. Correlation between the two cores was made by the present author. Width of cores not to scale.

100-130cm 130-145cm 145-195cm

195-225cm 10

See papers relative to the Harbour extension and the Government property in the French Creek (printed by order of His Excellency the Governor and laid on the table of the Council), Government of Malta, (1870: Plate II).

225-360cm

65

disturbed material; light yellowish brown (10YR 6/4) sandy silt; no material; yellowish brown (10YR 5/4 and 10YR 5/6) sandy silt with cobbles, iron staining at 50cm; brown (10YR 5/3) and pale brown (10YR 6/3) silty sand with cobbles; no material; clayey silt gradually turning from brown (10YR 5/3) to greyish brown (10YR 5/2) to grey (10YR 5/1), iron staining between 145-165cm; sand with gravel turning from greyish brown (2.5YR 5/2) to pale brown (10YR 6/3), iron staining at 205cm; grey (2.5YR 5/1) clayey silt with sand and scattered gravel. Oxide traces (iron

360-390cm 390-605cm 605-650cm 650-675cm 675-700cm 700-710cm 710-740cm 740-800cm 800-805cm 805-830cm 830-880cm 880-900cm 900-910cm 910-920cm 920-945cm 945-985cm 985-1055cm 1055-1070cm 1070-1120cm

staining?) between 250-280cm. Black streaks at 335cm. Smell of H2S noted by Frank Carroll. large boulder; grey (2.5YR 5/1) clayey silt with sand and scattered gravel and seashells; no material; large boulder; grey (2.5YR 5/1) clayey silt with sand and scattered gravel and seashells; large boulder; dark grey (10YR 4/1) silty sand with gravel; yellow (10YR 8/6) coarse sand; grey (2.5YR 6/1) silty sand with gravel and seashells, very wet; yellow (10YR 7/6) coarse sand; grey (2.5YR 5/1) very coarse sandy gravel; yellow (10YR 7/6) sand, very solidly packed; reddish yellow (7.5YR 6/6) sand with gravel; light yellowish brown (10YR 6/4) sand with gravel; strong brown (7.5YR 5/6) silty sand with gravel; reddish yellow (7.5YR 7/6) sand with gravel; dark red (2.5YR 4/6) and yellow red (5YR 5/6) silty clay with sand and gravel; red (2.5YR 5/6) silty clay; reddish brown (5YR 4/4) sandy silt. Solid bedrock starts at 1120cm, but large pieces of bedrock are present from 1075cm downwards;

625-650cm 650-680cm 680-700cm 700-715cm 715-755cm 755-785cm 785-805cm 805-820cm 820-940cm

4.4. Description of the Harrison & Company cores, Marsa Sports Ground The location of the boreholes made during the site investigation by Harrison and Company in 2000 is shown on Figure 4.1. At that time, the ground was still covered by asphalt and the general ground level at the site was 2.5m above sea level. Exact levels are marked out on the plan. As the coring done by Harrison was non-retentive, the data from the three boreholes are very sketchy: Harrison Core 1: 0-80cm asphalt and clay 80-180cm clay 180-880cm no description 880-890cm water encountered 980-1080cm soft rock 1080-1180cm fissure 20cm deep, 11.5m solid rock.

Marsa Core 2: During the process of dividing the samples for analyses, the ‘wet’ Munsell colour of every sample was determined, and general remarks about some of the sediments noted. As this core has not been processed beyond this point, there are no scientific particle size analysis data available. Thus, Frank Carroll’s remarks here form the basis of the sediment descriptions. The stratigraphic sequence for Marsa Core 2 was as follows: 0-30cm 30-80cm 80-95cm 95-120cm 120-170cm 170-220cm 220-255cm 255-385cm 385-520cm 520-590cm 590-625cm

dark brown (5YR 3/2) to brown (10YR 4/3); yellowish brown (10YR 5/4) to brown (10YR 5/3) sand with gravel; stone obstruction; reddish yellow (7.5YR 6/8); strong brown (7.5YR 5/6) sand with gravel decreasing; reddish yellow (7.5YR 6/6) sand with gravel increasing; strong brown (7.5YR 5/6 and 4/6); dark reddish brown (5YR 3/4); strong brown (7.5YR 5/6) gravelly with a dark reddish brown (5YR 3/3) streak between 865-870cm. A boulder obstructs between 875-880, the sediment becomes increasingly gravelly until bedrock is reached at 935cm.

Light yellowish brown (10 YR 6/4); no material; Yellowish brown (10YR 5/4); greyish brown (2.5YR 5/2); large boulder; grey (2.5YR 6/1 to 5/1) silty sand up to 190cm, then silty clay; no material (boulder?); grey (2.5YR 5/1), initially very stony (up to 270cm), then silty clay with sea shells; dark grey (2.5YR 4/1) silty clay with a sandy streak between 405-420cm; no material; dark grey (2.5YR 4/1) clayey; 66

Harrison Core 2 0-80cm 80-180cm 180-380cm 380-480cm 480-580cm 580-680cm 680-780cm 780-880cm 880-980cm 980-1080cm 1080-1180cm

asphalt and clay clay no information H2S smell hammer action ceased no information softer clay no information water encountered clay returns bedrock at 1150cm.

Harrison Core 3: 0-80cm 80-380cm 380-480cm 480-580cm 580-980cm 980-1080cm 1080-1180cm

asphalt and clay clay no information H2S smell no information soft rock fragments bedrock at 1120cm.

4.5. Results A number of interesting points result from the stratigraphies of Marsa Core 1 and Marsa Core 2. While it has hitherto only been presumed that the sea at Marsa extended further inland (e.g. Haslam & Borg, 1998: 108113; de Bono 1998: 2), the discovery of marine sediments (Figure 4.2. and see Appendix I) not only provides evidence for a former marine environment in the area covered by the present day Marsa Sports Club, but also shows that this environment prevailed for a considerable length of time. For Marsa Core 1 it emerges from Figure 4.3. that the deposition of the sediments was not a linear process, where a land borne sediment, possibly from the Late Pleistocene or Early Holocene, was superseded by marine sediments after the end of the last glaciation and throughout most of the Holocene, to be followed by alluvial deposits, as appears to have been the case for Marsa Core 2, as well as for example for Malta Shipbuilding Core No. 9 (Figure 4.4.). Instead, the Munsell ‘wet’ column in Appendix I shows considerable variations not only within the alluvial and marine deposits, but also alternations of these two different kinds of deposits. While the wet soil colour gives the impression that there were large layers of a homogenous deposit, the dry Munsell column highlights numerous fine variations within the various layers of Core 1. This may be indicative of different degrees of leaching during or immediately after the deposition of the sediment, but may also be the result of (bio-) chemical reactions within the sediment. There are no comparable data from Core 2, but the picture is likely to be similar.

Figure 4.7.: Part of Sample 170 of Marsa Core 1, at 980cm. The segment shows thin layers of different sedimentation events. Underlying milimeter paper for scale.

encountered on lake bottoms. The preservation of lamination at the bottom part of Marsa Core 1 may be due to waterlain sedimentation and a lack of bioturbation, as for example indicated by the presence of only 1 subterranean land snail species (at 1060cm) in the lower part of the core. Both Ceciliodes acicula or Hohenwartiana hohenwarti (see Appendix II) belong to the family Ferussaciidae and have been found living at depths of up to 2m below the ground surface (e.g. Evans, 1972: 168). They can actively burrow in friable soil, but can also follow root channels to these depths, their presence in other samples further up can thus perhaps give an indication of the ground surface level (Preece et al., 1998: 45). However, while in other parts of the core the Ferussaciidae are likely to have been washed into the deposit with erosive rainfall events, the good state of preservation of its friable shell in the basal part would speak against such an origin. The lamination broadly extends down to around 1075cm, although it must be borne in mind that the soil record is of a rather low resolution due to the method applied.11 The variations in the dry Munsell colour at the base of the core between 1120 and 1075 may be due to pedogenic processes and podzolisation. The moist soil samples here showed no signs of lamination (see Appendix I).

Taking also into account the limited data from the Harrison Cores 1, 2 and 3, it emerges that the bedrock level is quite irregular, despite the comparatively short distances between the cores (Figure 4.2.). The ‘water encountered’ in Harrison Cores 2 and 3 is likely to represent the lowermost marine layer of Marsa Core 1, as they appear roughly at the same distance below sea level. The smell of H2S, which was noted in Harrison Cores 2 and 3 was also noted by Frank Carroll during the processing of Marsa Core 1, at roughly the same distance below sea level. Neither the smell, nor a marine layer at the lower end of the core appear in Marsa Core 2, thus the relationship between the Harrison Cores and Marsa Core 1 appears much closer.

The appearance of a grey layer at 880cm in Core 1 indicates a sudden event that rapidly covered the underlying light yellowish brown sediments. That this event may have been violent is indicated by the presence of predominantly rounded gravel and sand. The grey colour of the sediment is likely to be due to reducing conditions because of waterlogging. Reducing or anoxic conditions occur when stagnant water becomes depleted of oxygen, and minerals are formed that contain iron in the reduced state Fe2+ (Evans & Heller, 2003: 158). The rounded gravel furthermore indicates the waterborne origin of this deposit. There appears to be no corresponding layer in any of the other cores under investigation, although perhaps the presence of water

4.6. Interpretation The accumulation and retention of red-brown coloured sediments at the bottom of both Marsa Core 1 and Marsa Core 2 may indicate that both cores are located in some sort of hollow, where the sediments are trapped and are not easily washed out by either freshwater runoff or seawater currents. Some of the samples were laminated, at times finely (e.g. Sample 170 at 980cm, see Figure 4.7.), but also coarsely (e.g. Sample 175 at 1005cm). This could indicate that the terra rossa sediment did not accumulate as a result of a single event, but in a series of events over a longer period of time as a result of erosive processes, not unlike the varved sediments usually

11 For comparison, cores containing lake sediments are usually cut lengthwise and/or x-rayed to reveal the laminations (e.g. Sadori et al.: 2004). Owing to technical limitations, this was not possible in our case.

67

noted in the coring report for Harrison Cores 2 and 3 could point to a continuation of this deposit on a small scale.

grey tone as those of the corresponding level in Marsa Core 1. The origin of the large boulders that occasionally occur in both cores is difficult to ascertain, especially since the material could not be retained during the drilling process. It is tempting to link the large boulder noted in Marsa Core 2 close to the mean sea level with anthropogenic activity, possibly some architectural feature,12 but the evidence is too sketchy to allow any conclusions.

The grey deposit in Marsa Core 1 also experienced a rapid burial by a 90cm thickness of nearly sterile yellow sand between 830cm and 740cm, which again was interrupted by a shallow marine deposit of ca. 5cm thickness at 800cm. This sedimentary sequence is not mirrored in Marsa Core 2, although the events leading to this sequence may have left a different signature there, as indicated by the accumulation of alternating strong brown and reddish yellow sediments at the same level (see also below). The discrepancy between the cores may perhaps suggest a channel feature in Marsa Core 1, and/or perhaps a buried terrace in Marsa Core 2.

In Marsa Core 1, the grey silty sediments gradually turn sandier and browner as present day sea level is approached, indicating that the area is slowly being filled with material from alluvial mudflows that also contains coarser particles. While the basin filled in, the coastline was thus pushed further towards the sea. This is also reflected in Marsa Core 2, where above the large stone boulder at around sea level the deposits turn from greyish brown to yellowish brown.

The calcrete layer in Marsa Core 1 probably created an impenetrable barrier between the overlying dark grey sediment at 740cm and the underlying strata. It is difficult to ascertain whether the calcrete formed in situ as its formation may take around 2000 years under fluctuating climatic conditions from semi-arid to sub-humid through the dissolution and precipitation of calcium carbonate in a freshwater vadose zone (MacLane, 1995: 319). It is thus also possible that the calcrete had formed elsewhere at a much earlier time and was deposited at the coring site of Marsa Core 1 as a clast.

4.7. Conclusion Owing to the absence of any C14 dates or any other comparable data from Marsa Core 2, it is not possible to securely relate the stratigraphy of Core 1 to the corresponding level in Core 2. Despite the relatively short distance between the cores, there are differences in the stratigraphical sequence resulting from differences in topography that perhaps also indicate channel migration. Such differences occur on a larger scale across the Marsa Basin, as implied by comparison of the Marsa Cores with those from Malta Shipbuilding and from Il-Menqa. Despite the absence of a rock contour record at the Marsa Sports Ground on a comparable level to the one in Figure 4.6. from Malta Shipbuilding, the latter helped to draw some conclusion on the underlying topography at what is today a plain on which the Marsa Sports Complex is sited.

The base of this dark grey sediment represents the beginning of a long sequence of waterlogged conditions in Marsa Core 1. At the corresponding level in Marsa Core 2, there are still yellowish brown to dark brown sandy deposits (see Figure 4.2.). As the dark grey colour of the sediment would point to anoxic or hypoxic still water, it would appear that, at the site of Marsa Core 1, there may have been a shallow water body severely depleted of oxygen, perhaps due to eutrophication. A large boulder covers this layer in Marsa Core 1 between 710cm and 700cm and there is a corresponding stone obstruction in Marsa Core 2. Whether these two occurrences are related or not is difficult to ascertain, as in Marsa Core 1 the boulder is overlain by grey deposits, while in Marsa Core 2 sandy yellowish and dark brown sediments follow. But although there are several stone obstructions in both cores, this appears to be the only occurrence at the same level. In Marsa Core 2, the dark brown sediments are abruptly followed by a dark grey deposit at 625cm. This may be due to rising sea levels. In Marsa Core 1, a grey silty clay with gravel and shells accumulates after another stone obstruction between 675cm and 650cm. Both cores, however, do not appear to have shared the same oxic conditions as indicated by the different grey tones, pointing to different levels of reduction. It appears that conditions at the location of Marsa Core 1 were more oxic, while they appear to have possibly been eutrophic for Marsa Core 2 as indicated by the dark grey coloured sediments. Oxic conditions improve in Marsa Core 2 at 385cm, and at this level, the sediments share the same

12 The recently discovered architectural feature during the widening of the stormwater canal at the Marsa Trade School showed that large ashlar blocks were resting directly on grey silts below sea level (A. Bonanno, personal communication, 6/2005)

68

overlying rubble, as can be still seen in the section in Figure 1.3. However, the recent discovery of a black and white aerial photograph of the Marsa area dating back to the early 1930s shows tree growth on the site (Figure 5.1).

Chapter V RADIOCARBON DATING OF MARSA CORE 1 5.1. Introduction Unlike the radiocarbon dates from the Maltese Islands discussed above, which all came from various archaeological contexts (see Chapter III), the dates reported hereunder all come from geological contexts. Thus, it is not archaeological material of a specific cultural phase that has been carbon dated, but the stratigraphical sequence of Marsa Core 1. The radiocarbon age determinations reported below will provide the chronological framework for the present study. All nine radiocarbon dates were determined by Beta Analytic Inc. in Florida, using different organic material found within the various sediments and levels of Marsa Core 1 (see below, Table 5.1.). Sample pretreatment consisted of alkali-acid-alkali washes, all done by the laboratory. Due to the overall low amount of datable organic material found, all dates were determined with the AMS (Accelerated Mass Spectrometry) technique. This method only requires very small amounts of organic material for the determination.

Figure 5.1.: Aerial photograph most likely taken in the early 1930s, showing the Marsa Sports Ground and race track in the foreground, and the Grand Harbour and inlet at Marsa behind. The red spot (arrow) marks the coring site, which was then covered by trees. Source: downloaded from Ebay, origin unknown.

Waterlogging of many meters thickness of sediments within the core and rapid burial provided anoxic conditions that helped to preserve many organic remains such as wood, plant fragments, seeds and bones. However, their amount and distribution varied considerably throughout the core, and sometimes a sample of possible interest contained no datable material at all. In this case, the closest sample with sufficient datable material was chosen. Two samples of interest, MRS 1-193 and MRS 1-149 contained too little macroscopic charcoal or other suitable organic remains for the AMS technique. As neither the neighbouring samples contained sufficient datable material, it was decided to date the bulk sediment. This involved the dating of microscopic charcoal, which was extracted chemically by Beta Analytic Inc from ca. 200g of sediment.

Analysis of the molluscan remains from a 3.52 m test core1 from the Marsa Sports Ground indicated that the sediments underlying those levelled by bulldozers showed a logical sequence of deposition, in the sense that the molluscs indicated how a formerly brackish water environment gradually silted up and turned into land. As the location of the coring site of Marsa Core 1 lies in what is today an alluvial plain, a large percentage of the sediments are allochthonous, but these may also contain autochthonous components. Hence, there are some difficulties associated with the choice of suitable organic material for dating. 1.

2.

3.

The waterlogged marine sediments showed no laminations and the faunal content (molluscs and crabs, see below), indicated bioturbation. Through this bioturbation, plant material and charcoal may also have been transported after deposition to different depths within the column. Several ‘events’ characterise the stratigraphical sequence. Among these, the transportation and subsequent deposition of vast amounts of sediments during floods is likely to contain material that may, depending on the size of the event, be considerably older than the event itself. There were no living trees overlying the coring site for several years prior to the site investigations, but several trees still surround the site today. Root penetration or action directly at the coring site was not taken into consideration as the ground surface of the site itself prior to the excavation works consisted of concrete

5.2. Results The conventional radiocarbon ages and 13C/12C ratios were determined by Beta Analytic Inc for nine samples from Marsa Core 1. The 2 sigma calibrations were made by the present writer using the most recent calibration software (OxCal v3.10, Reimer et al., 2004; Bronk Ramsey, 2005). The age-depth relationships of the dated samples (2 sigma calibrated) reveal several hiatuses (Figure 5.2.). Table 5.1. lists all details relevant details of the radiocarbon determinations. 5.3. Discussion As with all dates, the determinations were made on the assumption that the samples measured are related directly or indirectly to some event, the age of which is required in the investigation. Several observations may be made on the above results.

1 The core had been extracted in April 2002 with a hand soil auger for test purposes at a distance of 1 m from the location of Marsa Core 1.

69

1990). However, as the contaminated dates are of no use for the present study, they will not be considered further. At 48cm above the beginning of the core there is a layer of clinker. This layer was laid at around 1869 by the British to provide a firm bedding for the overlying race track. Although outside the core, this provides a suitable top anchor for a known date as its location can easily be linked with the core (compare Figures 1.4.1. – 3.). Furthermore, the sediments below the clinker layer that eventually link up with the core show no visible signs of disturbance (see Figure 1.4.3.). The 13C/12C ranges are all as expected (ibid.). The radiocarbon dates of MRS 1058, 1-090 and 1-120 all provided results within the expected ranges, as did MRS 1-162, were it not for the result of MRS 1-149 (see below). The date range of MRS 1-090 is unfortunately quite wide, and the 2-sigma calibration curve intercepts twice with the radiocarbon age and the calibrated curve time scale (see also Figure 5.3.). Nonetheless, Figure 5.2. indicates that the sedimentation rates were comparatively low between MRS 1-120 in the Neolithic Period and MRS 1-090 in the Phoenician/Punic Period and accelerated during the early Neolithic Period and from the Byzantine era onwards. Three dates (MRS 1-149, MRS 1-186 and MRS 1-193) appear excessively old (of Late Pleistocene age, MRS 1186 perhaps older2). Again, a mismeasurement appears unlikely because the associated 13C measurements all lie within the expected ranges (ibid.) and there appears no reason to distrust these dates more than any other date reported here. The difference of approximately 14,000 14C years between MRS 1-149 and the date that might be expected for that layer (judging by the dates of MRS 1120 and MRS 1-162) is of a magnitude highly unlikely to occur by chance (see Ashmore et al., 2000: 101). The same also applies to MRS 1-186, where the difference amounts to more than 24,000 14C years compared to MRS 1-193 and MRS 1-149.

Figure 5.2.: Radiocarbon dates of Marsa Core 1, determined by Beta Analytic Inc. Depth-age relationships for calibrated (2 sigma) radiocarbon dates, excluding the two contaminated dates (see Table 4.1.);

Two determined dates (MRS 1-007 and MRS 1-108) appear to be contaminated with modern carbon. The material used for both dates were similar appearing plant fragments, which could not be defined further. As both samples were located at the top end of a sample tube, it is possible that the contamination occurred when the tube was pushed into the sleeve and may have picked up organic material from the surface. A possible mismeasurement by Beta Analytic seems unlikely, as the laboratory ran MRS 1-108 again and obtained the same result. Furthermore, the associated 13C measurements are within the expected ranges and would also point to an exact measurement (Gupta & Polach, 1985: 114; Aitken,

These ‘old’ dates may be the result of a variety of reasons. In the case of MRS 1-149 and MRS 1-193, the total organic content contained in a bulk sediment sample was dated and not an individual fragment or piece of

Sample No

Lab Code

Depth (cm)

Material

13C/12C ratio

MRS 1-007

Beta - 208958

50

plant fragments -26.2 o/oo

Conventional Radio- 1 σ calibrated result carbon Age (68.2% probability) 106.8 +/- 0.4 pMC 1955 AD

2 σ calibrated result (95.4% probability) 1952 AD

MRS 1-058

Beta - 208961

320

plant fragments -25.6 o/oo

1460 +/- 40 BP

570 AD – 640 AD

MRS 1-090

Beta - 200517

510

Charcoal

2510 +/- 40 BP

780 BC – 540 BC

800 BC - 420 BC 1952 AD

-23.7 o/oo

530 AD - 660 AD

MRS 1-108

Beta - 208960

600

plant fragments -24.6 o/oo

107.3 +/- 0.4 pMC

1955 AD

MRS 1-120

Beta - 200518

730

Charcoal

-24.1 o/oo

5730 +/- 40 BP

4660 BC – 4500 BC

4690 BC - 4460 BC

MRS 1-149

Beta - 203318

875

bulk sediment

-23.8 o/oo

19530 +/- 90 BP

21600 BC – 21050 BC

21750 BC - 20800 BC

MRS 1-162

Beta - 200519

940

Seeds

-25.6 o/oo

5870 +/- 40 BP

4790 BC – 4700 BC

4840 BC - 4610 BC

MRS 1-186

Beta - 208959

1060

Charcoal

-23.7 o/oo

> 44400 BP

date out of range

date out of range

20020 +/- 90 BP

22180 BC – 21880 BC

22350 BC - 21750 BC

MRS 1-193

Beta - 203320

1095

bulk sediment

-24.3 o/oo

Table 5.1.: Details of the AMS radiocarbon dates from Marsa Core 1, determined by Beta Laboratories and calibrated with OxCal v3.10. For the contaminated samples the percentage of modern carbon is given.

2

70

The age of MRS 1-186 lies beyond the measurable radiocarbon age.

charcoal. It emerges that the total organic content contained an unknown percentage of ‘old’ recycled carbon. The overall generally very high sedimentation rates between 730cm and 940cm (average of 5 cm every 3.81 years) would suggest that significant transport of ‘old’ sediments by geomorphological processes such as flooding may be responsible for the age-depth anomalies. The age of MRS 1-186 was determined from a single piece of charcoal that had the size of a pin head. Although its determined age lies beyond the measurable limit of radiocarbon, its presence also may indicate the allochthonous nature of the deposit. Episodic events that involve rapid sediment movement of superficial deposits from adjacent hillslopes have been detected in only few studies so far, as this generally needs to be supported by a detailed dating framework (e.g. Ashmore et al., 2000; Stanley & Hait, 2000; Warburton et al., 2004; Bourke & Thorpe, 2005). While several episodic events in Marsa Core 1 could also be detected through the stratigraphical investigations (see below), the results of the carbon dating provided a vital time-frame that firmly locates in time rapid mass-movements particularly into the lower third of the core.

71

different beds vary in age,2 Trechmann ascribes great antiquity to them on account of their isolated patchy occurrence, which he attributes to later erosion. He also proposed the idea that the reddened and blackened pebbles that are often found within Quaternary deposits, may be the result of a scorching and arid period (ibid.: 1).

Chapter VI SEDIMENTOLOGICAL INVESTIGATIONS 6.1. Introduction Both Marsa Core 1 and Marsa Core 2 dug from the Marsa Sports Ground revealed several sediments ranging in radiocarbon age from late-glacial to recent times. Particle size analyses, water content, organic matter content, geochemistry and magnetic susceptibility analyses1 were employed to extract environmental information from Marsa Core 1. A mineralogical study to determine the sources of the major sedimentary units, as has been done in some other studies (e.g. Catt & Staines, 1998: 69ff) did not prove useful in this case as there are too few differences between the mineral composition of the various parent materials from which the sediments were derived. For example, following Murray’s mineral analysis of the Maltese geology (1890), only the presence of rutile in a sediment would point to a source area underlain by Globigerina limestone, while only the presence of magnetite would indicate sediment derived from the erosion of a greensand deposit (ibid.). Furthermore, in many cases, the crystal structure of the mineral grain would also need to be determined since several minerals share the same chemical formula, but differ in crystal structure (e.g. haematite (α Fe2O3) and maghaemite (γ Fe2O3). and similarly goethite (α FeOOH) and lepidocrocite (γ FeOOH). Such determinations were beyond the means and scope of the present study.

Since Trechmann’s study, several other types of Quaternary deposits were identified (e.g. at Hamrun in 1941 (MAR, 1946-47: IX) and at Mriehel in 1965 (MAR, 1965: 8-10)), but it was primarily the biological inclusions rather than the mineral components and composition of the sediment that were the focus of attention. Already in 1920, Dawson Shepherd laments the non-existence of a proper soil survey, when he formulated a report on agriculture in Malta for the government of the day (Vella, 2003: 235). The history of general soil studies in the Maltese Islands starts in 195657 when D.M. Lang made the first soil survey of the Maltese Islands, following Kubiena’s 1953 soil classification system; Lang’s report and soils map was published in 1961. The aim was to provide basic descriptions of the soils and to map their distribution as an aid to agricultural planning. Lang’s soils map was based on detailed field observations of a total area of ca. 30 km², while the remainder was mapped from aerial photographs and several ground-truthing observations along traverses (Vella, ibid.). Although the different soils and series were generally described from top soil level to parent material, the map outlines only the visible soils and soil complexes (Lang, 1961: 86-93). While Alcol series deposits were recognised in the Burmarrad plain at Salina, the soil data for the Marsa plain were little more detailed than noting the presence of Xerorendzina soil (ibid., 84-85). The Malta Structure Plan (Planning Authority, 1990) still relied on Lang’s findings and the State of the Environment Report (MEPA, 1998) highlighted the need to survey the soil resources anew (MEPA, 2002b: 313), and to this end the MALSIS project was started in 2002. This most recent soil survey and soil mapping of the Maltese Islands also incorporated the deep colluvial soils found in some of the broad ancient valleys on the island (Vella, 2003: 238), but this time, there are no data at all for the Marsa plain. The reason for this is that the MALSIS project was concerned with agricultural soils and the Marsa plain is on one hand industrially developed, and on the other occupied by the Marsa Sports Ground. The part of the Marsa Sports Ground, where the coring for the present study took place, has been scheduled for being ‘an important geomorphological unit, also of scientific significance in hydrology, geomorphology and Quaternary palaeontology’ (MEPA, 2002a: 75), without giving any further information. Hence, there are very little comparative data available for the cores from the Marsa plain.

6.2. History of sediment studies in the Maltese Islands Early soil studies of the Maltese Islands concentrated mainly on the description and investigation of Quaternary deposits (Adams, 1870; Feilden & Maxwell, 1874; Cooke, 1891). This is because these deposits formed during the Pleistocene under different climatic conditions, which have since been superseded (Lang, 1961: 83) and as such often contain the remains of extinct animals or past ecological assemblages (e.g. the Tal-Gnejn fissure near Mqabba, the Maghlaq deposit near Mnajdra and the Bahrija Quaternary deposit). The Ghar Dalam cave and deposits highlighted particularly well the changing faunal assemblages and thus received considerable attention (e.g. by Despott from 1917 onwards and by Baldacchino in the MARs between 1933-1938). Trechmann (1938: 1-26) then classified the Pleistocene beds of the Maltese Islands into three categories: valley loams and breccias, which are often partly washed out again by streams or by the sea; conglomerates and breccias that border or are near the coast, often thinning out as calcareous veneers; and ossiferous deposits in caves and fissures. Although these

1

2 The bones of the various animals within these deposits give an indication of their age as faunal assemblages varied considerably between the different glacial periods (see below and Hunt & Schembri, 1999: 41-75).

Frank Carroll conducted the analyses of every fifth sample (i.e. Samples 1, 5, 10, 15 etc.) for organic carbon content, geochemistry and volume and mass specific magnetic susceptibility. I am very grateful to him for making his results available for the present study.

72

calculate the dry weight of the whole portion3 (by simple proportion). The latter is very important to standardise the weight of sediment as all portions had a different volume and water content and thus needed to be standardised for the quantitative comparisons of the biological components.

6.3. Particle size distribution The main objectives in determining the particle size distribution of a sediment are description, comparison and interpretation. The particle size is a fundamental physical property of deposits and reflects weathering and erosion processes that generate grains of different sizes, and the nature of subsequent transport processes (Bloggs, 2001: 59). In aquatic deposits, particle sizes are also closely linked to turbulence, wave energy and proximity to the shoreline. An increased grain size is generally linked with higher energy conditions of sediment production or transport, whereas lower energies are indicated by smaller particle sizes. However, an increase in sand- and larger sizes through time may also reflect low water levels, related to a drier and warmer climate (Bowerman, 2003: 39).

The samples then needed to be broken up to separate, with the least possible damage, the lithic and soil and clay particles from the biological components. For this process they were soaked in a solution of sodium hexametaphosphate4 in glass containers for as long as necessary. Very clayey samples were additionally treated with hydrogen peroxide to facilitate the process. Once the break-up process was completed, the samples were wetsieved through nested 500 micron and 63 micron sieves. Particles smaller than 63 microns (silt and clay) were discarded. After drying in the oven at 70º C, the >500 micron section was divided into a sub-fraction of particles larger than 8mm (medium pebbles, gravel and stones) and a sub-fraction of particles larger than 500 microns (very fine to medium sand) but smaller than 8 mm (coarse sand to fine pebbles). These two sub-fractions and the smaller than 63 micron fraction (very fine to medium sand) were then weighed again to establish the granulometric profile. The sum of the three fractions per sample was then deducted from the calculated overall dry weight of each sample to establish the weight of the discarded silt/clay section ( 8m m

> 500m ic

> 63m ic

< 63m ic

(%)

(% )

(%)

(%)

(% )

(%)

10

20

30

40 50 0

1

2

3

4

5 0

20

40

60

80

0

10

20

30

40 50 0 10

20

30

40

50

0

20

40

60

tota l w e ight (g) 80

0 100 200 300 400

0

1

2

3

depth (m)

4

5

6

7

8

9

10

11 PANGAEA/PanPlot

max.: 11.15 m

section components1.txt - 03.07.2006 10:30 h

Figure 6.1.: Water content, organic matter and particle size distribution throughout Marsa Core 1. Source of organic matter data: Frank Carroll

74

A sudden increase of sand and small pebbles is combined with a sharp decrease in the silt/clay fraction at 220cm. Although the silt/clay fraction recovers slightly at 195cm, there is a marked general trend for an overall increase in the sand fraction combined with a general decrease in the silt/clay fraction from 220cm upwards. From 120cm, which corresponds roughly to the present day sea level, upwards to 50cm, the mud content steadily increases, while the sand and gravel continually decreases.

somewhat surprising, as Terra soils have the highest organic matter content of all Maltese soils: between 3.1 and 6.4% (Lang, 1961: 88). However, there is a direct relationship between the accumulation of organic matter and fine-grained muddy sediments, as organic matter adsorbs onto mineral surfaces. This process helps to preserve the organic matter as it delays decomposition (Hedges & Keil, 1995: 81). The aquatic sediments between 695cm and 225cm that have a high silt/clay content, illuminate this particularly well (see Figure 6.1.). Here, the organic content reaches its peak of 4.37 % at 435cm, while a low organic content is generally associated with a sandy sediment. That the overall level of organic content reaches only 4.37% may also be due to low sedimentation rates. These allow a longer contact time between organic matter and dissolved oxygen in the water column, which therefore can contribute to decreased concentrations of carbon and nutrients in sediments through degradation (Hartnett et al., 1998: 572). Furthermore, when organic matter is degraded in waterlogged conditions by aerobic bacteria, the dissolved oxygen concentrations are usually lowered and anoxic and hypoxic conditions may develop (ibid.: 574). As the aquatic sediments in Marsa Core 1 are generally grey to dark grey, this suggests anaerobic decomposition process acting on the organic matter. On the other hand, enhanced sediment transport by wind and water as a result of erosion can lower the organic content of a sediment because the minerals and clays dilute its concentration (Radke, 2002: 6). This may possibly be observed in the upper 220cm of the core, where the organic matter content hovers around 3% as the basin silts up.

The water content also varies greatly throughout the core. Similar to the poorly sorted sediments between bedrock and 700cm, the water content follows no detectable trend, apart from being generally below 20%. The few peaks in this section appear to be correlated to overlying coarse sediments, which, due to their pore capacity, enable better water percolation, while a high mud content at the peaks would favour the retention of percolated water (see Figure 6.1.). On the other hand, coarse sediments with a low mud content show a markedly low water content, which could indicate that the water drained downwards. This trend may also be observed in the overlying aquatic sediments between 695cm and 220cm, although here the average water content is higher at between 25% and 35%. This is possibly due to waterlogged conditions, which prevent water loss through evaporation, and the clastic nature of the silts and clays. The highest peak, at 390cm with a water content of 45%, is perhaps the result of the overlying boulder, which, through on the pressure that it exerted on the underlying substratum, would have forced extra water into the muddy sediment below. Between 195cm and 50cm, the water content again lies below 25%. Since the sediments above 120cm lie above present day sea level, the low water content may be here be due to evaporation.

6.5. X-ray fluorescence (XRF) analysis – major and minor elements by quantity. The elements present in a sample can be identified by measuring fluorescent X-ray energies and comparing their values with standard measurements for each of the elements. The analysis is based on the excitation of the inner electrons of the atom, whereby the sample is irradiated with a beam of X-rays that excite electrons in the inner shells of all atoms in ground compressed samples. This causes the electrons to move up to a higher energy state, from where they immediately revert back to their original state, emitting in the process specific amounts of energies in the form of characteristic X-ray frequencies (Renfrew & Bahn, 1996: 344-345). The instrumental limit of detection of the various elements may be very low (at around 1 to 2 parts per million, depending on the element), and although this value has been found to be much lower than the smallest concentration of the analyte that can actually be determined by the instrument (Rousseau, 2001: 37), it appears that for trace elements, this determination limit was always exceeded (see for example Cadmium (Cd) levels in Figure 6.2.).

6.4. Organic content The organic content was calculated by Frank Carroll, who followed the weight loss on ignition method described by Gale & Hoare (1991: 263-4). Here, a dried sub-sample of around 5g, generally from every fifth sample from Marsa Core 15 was weighed and placed into pre-weighed porcelain crucibles and put into a furnace at 450º C for 24 hours. The percentage of organic matter was then calculated from the weight loss that occurred between the net dried sample weight (i.e. without the crucible) and the net fired sample weight (also without the crucible) using the following formula: (dried – 450º fired sample weight) x 100 = % organic matter dried sample weight

The organic matter content was found to be generally low, between 1% and 4.37% only. It is comparatively lower in the lower part of the core, between 750cm and bedrock. Considering that this part supposedly contains Terra soils, the low carbon content would appear

Geochemical data may be very useful to establish human impact due to industrial activity (e.g. Carroll, 2001; Pyatt et al., 2000; Le Roux et al., 2003), but also to investigate the extent of chemical decomposition and weathering of sediments (e.g. Singh et al., 2003; Jarvis et al., 1998) and

5 Frank Carroll analysed Samples 1 and 5, and from then onwards every 5th sample, except for Sample 115. Here, the sample consisted of a solid rock without sufficient soft sediment material to conduct any analyses.

75

Fe 0

10000 20000 30000 40000

Cu 0

10

20

Zn 30

40

50 0

100

Cd 200

300 0 2 3 4 5 6 7 8 9

Pb

11 13 15 17 19 0

10

20

30

40

50

0

1

2

3

depth

4

5

6

7

8

9

10

11 PANGAEA/PanPlo

max.: 11.15

Revised Marsa geochem1in5.txt - 03.07.2006 08:40 h

Figure 6.2.: XRF determined metal components in Marsa Core 1, expressed in μg/g for every fifth sample. Source of data: Frank Carroll.

to determine climatic and environmental conditions (e.g. Grattan & Gilbertson, 1999; Calanchi et al. 1996) through changes in the concentrations of selected elements in a continuous record. The sources of these elements may vary, which often does not allow one to make a straightforward interpretation (ibid.: 85). Furthermore, as the broad chemical compositions of the different soils in the Maltese Islands are very similar (see above, Chapter II), it is of little use for determining the provenance of the sediment (see Canti, 1995: 185).

published a percentage breakdown of the major constituents of the different soil types,6 of which only calcium (Ca) and potassium (K) have been detected by the XRF analysis. Therefore, interpretation of the results generated by Frank Carroll should be regarded with caution. Yet, even in the absence of detailed background data, significant variations in the abundance of the various elements may give indications of changing environments and impact episodes. Figure 6.3. shows a strong similarity in the profiles for Aluminium (Al) and Silicon (Si), the major components of clay, throughout the length of Marsa Core 1, and there is a parallel similarity in the profiles of Potassium (K) and Titanium (Ti). These elements form part of the lithic environment of the Maltese Islands, while Calcium (Ca) is the main component of limestone and limestone derived sediments (Lang, 1961: 88). The latter is, therefore, the most common element detected here. There is a negative correlation between Al, Si and Ca, which is particularly apparent between the 680cm and 1080cm levels of the core. Here, the steady and marked increase of Ca towards 735cm combined with a steady decrease of Al and Si appears to reflect the calcification of the sediment, which culminates in the calcrete deposit found at 735cm.

The geochemistry of Marsa Core 1 was determined by Frank Carroll using the University of Huddersfield School of Applied Sciences’ XRF (X-Ray Fluorescence) Spectro X-Lab 2000 instrument, on ground compressed samples and is expressed by quantity in μg/g for the major and minor elements. Low atomic number elements were not determined by the XRF analyses conducted, and the determinations encompass several elements from Aluminium (atomic number 13) to Uranium (atomic number 92). As not all elements detected by the XRF analysis are relevant to the present study, only a selected data will be used here. 6.5.1. Results All elements detected and determined by the XRF analyses occur naturally in the earth’s crust and/or form an integral part of biogeochemical cycles. However, there are no published data so far of the average background trace element levels for any sediment of the Maltese Islands. With regards to Maltese soils, Lang (1961: 88)

Manganese (Mn) may be used to indicate mechanical or chemical weathering. Mechanical erosion is implied 6

These consisted of calcium (Ca), magnesium (Mg), sodium (Na) and potassium (K) only.

76

Al 0

20000 40000

Si 0 40000 90000

Ca 0

200000

K

Ti

0 5000 11000 17000 0

2000

Fe 4000

0

20000

Mn 40000

0

100

200

Mn/Fe 300

0.00

0.01

0.02

0

1

2

3

depth

4

5

6

7

8

9

10

11 PANGAEA/PanPlot

max.: 11.15

SI-3.txt - 03.07.2006 08:55 h

Figure 6.3.: XRF determinations of major elements in Marsa Core 1 for every fifth sample, amounts expressed in μg/g. Source of data: Frank Carroll.

where Mn dominates over iron (Fe), while the chemical weathering of soil profiles is suggested when the Mn/Fe ratio changes to favour iron (Grattan & Gilbertson, 1999: 85). In Marsa Core 1, the ratio appears to imply a permanent mechanical erosion, which seems to start off comparatively low (below 0.01) at the base of the core, but increases steadily to peak at 735cm. This peak coincides with the deposition of the nearly sterile sandy sediment, which turned into calcrete. The ratio decreases drastically with the waterlogging of the coring site through rising sea levels and the ratio stays below 0.01 between 680cm and 230cm. This may imply perhaps relatively steady sedimentation in the newly formed lagoon, although the sediments are derived from erosion. Further up, at 205cm and towards the top of the core, the ratio increases and the peaks coincide with a strong increase in sand-sized and larger particles, which possibly indicates the occurrence of an erosive event that led to the silting up of the area (compare Figure 6.1.).

available. Compared to sediments in Southern Jordan, copper (Cu) and lead (Pb) levels appear to fit very well within the ranges determined there for an uncontaminated context (Pyatt et al., 2000: 773; see also Le Roux et al., 2003: 741 for comparative data from Lindow Moss). Both Cu and Pb have several peaks throughout, but their occurrence cannot be linked with some industrial activity. Cadmium (Cd) was detected and determined throughout the core in levels that are well within the range for sedimentary rock7 and there is a prominent peak at 735cm. This peak coincides with a peak in Cu and Pb levels and may be related to the strong erosion and rapid deposition event that occurs at this level (see Cook & Morrow, 1995). The normal background concentration in sediments for Zinc (Zn) is around 90 ppm (Manheim, 1999), and the sediments from Marsa Core 1 generally fall well into this range. Even the peak at 305cm lies below an acute toxicity level (ibid.). Zn appears to behave antagonistically to Fe in the lower part of the core (see Figure 6.2.). This disagreement may indicate a lack of mixing of the sediments as they are deposited on the coring site of Marsa Core 1 (Kimball, 1997). The absence of high Zn levels coupled with a similar prominent peak in either Cu and/or Pb anywhere in the core may be perhaps be indicative of an absence of anthropogenic industrial activity in the harbour area.

Sulphur is primarily associated with biological productivity in waterlogged sediments (ibid.) and waterlogged conditions are reflected by clearly raised sulphur (S) levels between 830cm and 855cm, as well as from 685cm up to 215cm (see Figure 6.4.). Here, the grey colour of the sediments also indicates reduced oxygen conditions under water (see above, Chapter IV). Indications for anthropogenic pollution in form of metal contamination are not easily identified from Figure 6.2. as there are no comparable data from the Maltese Islands

7

The established range lies between 0.1 and 25 ppm (Cook & Morrow, 1995), while the highest amount in Marsa Core 1 for Cd was 15 ppm (see Figure 6.2.).

77

manganese oxides are the main carriers of magnetism. The occurrence of these magnetic minerals is nearubiquitous in the natural environment, but as they are highly sensitive to changes in environmental conditions, they may be very useful palaeo-environmental indicators (Gale & Hoare, 1991: 201-2).

6.6. Magnetic susceptibility Magnetic susceptibility measurements are widely applied non-destructive and cost effective methods to determine the presence of iron-bearing minerals within sediments. Sediment samples of a known volume and a known weight are exposed to an external magnetic field, which causes the sediments to become magnetised according to the amount and kind of iron and manganese8 compounds contained within the sample.

The magnetic properties of till from homogenous source areas are similar to those of the bedrock from which they were derived, but this may not be the case where chemical or bio-chemical weathering has occurred (ibid.: 208-9). Erosion, transport and deposition also alter magnetic properties (Alvisi & Vigliotti, 1996: 285), as does waterlogging (Bettis III, 1992: 121) and the influence of fire (Church, 2005: 47). This is mainly due to the nature of iron, which very readily forms different compounds with oxygen and sulphur, depending on the supply of oxygen and/or sulphur. Oxidation of iron from Fe2+ to Fe3+ is a common weathering phenomenon in alluvial deposits, where the oxygen supply is high and/or the biological oxygen demand is low. Generally, oxidised sediments occur where the water table is or has been below the oxidised zone (Bettis III, 1992: 121). Reduction to the highly mobile Fe2+ occurs under waterlogging when the biological demand exceeds the amount of oxygen available through bacterial decomposition of organic detritus. Anaerobic sulphur bacteria can then respire the oxygen present in sulphate

All materials show some reaction when they are exposed to a magnetic field, though this reaction will be very weak in the case of conventionally ‘non-magnetic’ materials. Magnetic materials are referred to as ferri- or ferromagnetic. They retain their magnetic properties even beyond the exposure to a magnetic field. Conventionally ‘non-magnetic’ materials do not retain any magnetism after being exposed to a magnetic field. These materials are either paramagnetic (e.g. pyrite, iron carbonates, manganese carbonates) or diamagnetic (e.g. quartz, calcite, water), depending on whether they are respectively either pulled into or pushed out of the regions of a strong magnetic field (Thompson & Oldfield, 1986: 3). Generally, susceptibility depends upon the concentrations of strongly magnetic grains in a material (Jordanova et al., 2001: 1137). Iron oxides, iron sulphides and

kfd

κfd(%)(%) -10

0

10

χlf Xfd

Fe Fe

(10(µm3kg ?6m³kg - 1) ?¹) 20 -0.1 -0.0

0.1

0.2

0.3 0 10000

30000

50000 0

100

Mn Mn

S

200

300 400 0 4000 8000 12000 17000

Ca

S 0

Ca

100000 200000 300000

0

1

2

3

depth (m)

4

5

6

7

8

9

10

11 PANGAEA/PanPlot

max.: 11.15 m

Figure SI-4.txt - 03.07.2006 09:37 h

Figure 6.4.: Volume frequency dependent and mass specific magnetic susceptibility and some XRF determined elements (expressed in μg/g) for every fifth sample, Marsa Core 1. Source of data: Frank Carroll.

8 Although nickel and cobalt also have magnetic properties, they are only present as trace elements in soils. Thus, their tiny quantities are negligible compared to those of iron and manganese.

78

molecules. This liberates sulphides with which these mobile iron molecules then react to form FeS or FeS2 (pyrite) (Barnes, 1994: 20). These diagenetic processes that occur after the deposition thus reflect the physical changes in the depositional environment (Alvisi & Vigliotti, 1996: 285).

6.6.2. Results The presence of enviromagnetic minerals in Marsa Core 1 may be due to two major causes: at the base of the core (between 1120 cm and 1070 cm) they may have been created or transformed in situ by chemical processes, and possibly also by biologically mediated ones, because of the extremely high biological and chemical reactivity of iron (ibid.: 84). Above 1070 cm, these minerals would have existed elsewhere and would have been transported by wind, floods and sea currents with the sediments to the location.

6.6.1. Material and methods The mass specific magnetic susceptibility (χ) and the volume specific frequency dependence (κfd %) were determined by Frank Carroll for samples 1 and 5, and henceforth every fifth sample, using a Bartington MS2 meter and MS2 laboratory coil at Huddersfield University. Carroll followed the method described in Gale & Hoare (1991: 223-225) and in the Bartington MS2 manual, and exposed the known mass or volume of the samples to a dual frequency of 0.46 (low frequency lf) and 4.6 kHz (high frequency hf) to determine the mass magnetic susceptibility and volume specific frequency dependent susceptibility. Both parameters measure the extent to which susceptibility varies with the frequency of the applied magnetic field and the volume specific frequency dependent is calculated with the following formula:

Figure 6.4. shows the levels and variations of the different magnetic susceptibility measurements. The volume frequency dependence (κfd), expressed in %, is fairly uniform above 8.2% between the base and 880 cm, but plummets to around 5% between 855 and 830, rises again sharply afterwards, to decrease again to levels below 5% up to 115cm, interrupted by a sharp peak above 10% at 205cm. Towards the top of the core, the frequency dependence increases steadily to reach a maximum of 12,28% at 20 cm. There appears to be a strong correlation between the percentage and type of deposit: the waterlogged gleyed deposits all have levels at times well below 5%. This is likely to be due to the reducing conditions, where the finest magnetic grains are most easily dissolved (Gale & Hoare, 1991: 214). The red soil deposits at the base of the core have relatively uniform levels at around 9%. The peak at 205 cm, which is located around 80cm below pdsl, may indicate a sudden land-derived sedimentation, where, due to rapid burial, the magnetic properties of the minerals present could be retained (Evans & Heller, 2003: 103). Generally, the overall levels are well below 15%, the maximum recorded here was 12.28% at 205 cm.

κfd = [(κlf – κhf)/κlf] x 100. The volume susceptibility is defined by the relation κ = M/H, where M is the volume magnetisation induced in the susceptibility sample κ, by an applied field, H. Following this definition as described in Thompson & Oldfield (1986:25), volume susceptibility is a dimensionless quantity, expressed in SI values.9 To obtain the mass susceptibility χ, one has to divide the volume susceptibility (κ) by the density (ρ). As κ is dimensionless, χ has units of reciprocal density, expressed in m³/kg (Evans & Heller, 2003: 9). Thus, the applicable formula here is:

The mass specific susceptibility χ shows a broad correlation of peaks with the κfd between the base of Marsa Core 1 and 580 cm, (see Figure 6.4.). Here, the susceptibility also appears to be related to the kind of deposit: the values at the base are considerably higher than in the waterlogged deposits between 735 cm to around pdsl at 115 cm, where there is generally very little variation. Above that, towards the top of the core, the susceptibility again increases slightly, possibly because of drying out of the sediment and increased oxygenated conditions (Thompson & Oldfield, 1986: 75).

χ = κ/ρ The susceptibility records the presence of fine magnetic grains at the stable single-domain/superparamagnetic boundary, which lies at approximately 0.05micrometer. These grains are commonly produced during pedogenesis (Sharp & Dowdeswell, 1998: 90) and a high frequency dependence indicates enrichment of superparamagnetic particles, particularly in palaeosols (Evans & Heller, 2003: 71).

The considerable variations of the susceptibility at the lowermost 4m of the core may possibly be indicative of unstable conditions that perhaps include erosive events and possibly also fires. Fire produces secondary magnetic minerals in the soil by converting paramagnetic or antiferromagneitc iron compounds into mainly ferrimagnetic oxides (Thompson & Oldfield, 1986: 119), which would then lead to an increase in susceptibility values. The enhanced susceptibility may also be due to downslope enrichment through erosive processes (Thompson & Oldfield, 1986: 82). Both the high values of χ and κfd point to the presence of fine-grained secondary ferromagnetic minerals in surface horizons,

In the Maltese context, it is not the ferromagnetic minerals alone that determine the bulk magnetic properties. Due to the limestone derived soils, diamagnetic components such as quartz and, particularly, calcium carbonate, are magnetically extremely significant, as this not only leads to considerably lower values when compared with other studies (e.g. Jordanova et al., 2001; Church, 2005), but often also to less perceptible variations. 9

SI is an international system of units and stands for Système Internationale.

79

and the values from Marsa Core 1 compare remarkably well with those found by Oldfield et al. (1985) in a catchment in Maryland, USA (Gale & Hoare, 1991: 213). The lower susceptibility values between 955 cm and 930 cm may possibly point to an increased climatic aridity (ibid.), while the even lower susceptibility values between 850 cm and 830 cm coincide again with the presence of a waterlogged and gleyed deposit, as also further up, between 735 cm and 115 cm. Here, due to the reducing conditions that led to the dissolution of primary and secondary ferromagnetic minerals, the susceptibility is low, as are the κfd values discussed above.

with other magnetic effects (Thompson & Oldfield, 1986: 3). This diamagnetic force is pushed into the background from 735 cm upwards by other processes. 6.7. Discussion The basal deposits, as found at the base of both Marsa Core 1 and Marsa Core 2, have been found to contain Pleistocene elements unrecorded for this area.10 Sedimentological investigation indicates that the lowermost deposit at Marsa Core 1, between 1120cm and 1070cm, may be a soil that formed in situ, as possibly indicated by the chemical weathering marks on the adjoining bedrock (see the high magnetic susceptibility values (Figure 6.4.). General bedrock contour information for the area is unfortunately missing, and even when compared with the rock contour data from the Malta Shipbuilding studies, in the absence of physically comparing the dark brown soils from both sites, conclusions on the type of Terra soil prevalent at Marsa Core 1 should only be drawn with caution. Considering the dark colour of the lowermost sediment, it is likely to have formed between lapiés. This is also indicated by the presence of fragments of xerophytic plants (Lang, 1961: 92-93 and see below). Immediately overlying this deposit are deep brown to yellowish red soils, which are irregularly laminated. This is indicative of erosive processes, which would have led to the layered sedimentation as from 1070 cm, which may possibly be described as an Alcol series. According to Lang, this soil type is developed on the valley loams, (both Quaternary and recent), which are partly alluvial and partly colluvial, hence the name Alcol. Occurring on the broader valley bottoms, they are the erosion products of Terra, Rendzina and Carbonate Raw soils (ibid.: 91). In the case of Marsa Core 1, the deposits appear to be the erosion products of Terra soils, due to their striking reddish colour and also due to the presence of several teeth of the extinct Pleistocene dormouse Pitymys melitensis (see below, Chapter VII). The carbon dates determined by Beta Analytic Laboratories (Florida) for Samples 193 and 186 would corroborate this interpretation (see above, Chapter V). These erosion-derived deposits extend up to 880 cm and leave a signature that is very irregularly laminated. The absence of background data on Maltese soils with regards to their geochemical composition and magnetic susceptibility properties make interpretation of the present data somewhat hazardous, yet these data provide important indications.

Despite the presence of fine-grained mud, predominantly in the waterlogged deposits of Marsa Core 1, there is no apparent relationship between the particle-size distribution and magnetic susceptibility (compare Figure 6.1.), although studies elsewhere have demonstrated a close relationship between fine sediment particles and enhanced susceptibility (e.g. Gale & Hoare, 1991: 204205). As possible post-depositional transformation of the iron minerals (e.g. through fire, burial, exposure or waterlogging) may also influence the (bio-) chemistry and sedimentary occurrence of other elements such as phosphorus and sulphur in the sediment, the magnetic susceptibility data may be compared with the XRF data of some chemical elements of Marsa Core 1. Comparing both the χ and κfd results with some of the results from the geochemical analyses, there is no apparent relationship between the susceptibility values and iron content (Figure 6.4.). Increases or decreases of iron appear not to have any effect on the susceptibility. Manganese (Mn), the only other magnetic element measured, occurs in such small quantities that it shows no correlation with the susceptibility either. Similarly, there appears to be no relationship with the amounts and variations of phosphorus (P). Very interesting, on the other hand, are the sulphur (S) levels. As mentioned above, the raised sulphur levels indicate waterlogged sediments. This sub-oxic environment favours the diagenesis of iron and sulphur. While hydrogen sulphide (H2S) is produced by bacterial sulphate reduction, the dissolution of magnetic particles begins as soon as the boundary from iron rich Fe2+ > H2S to the sulphur rich system H2S > Fe2+ conditions is crossed in the anoxic zone. The persistently high sulphide concentration then leads to the dissolution, sulphidisation and finally the pyritisation of magnetite. The end product, pyrite (FeS2) is paramagnetic, and may be one reason for the low susceptibility values in the gleyed deposits. Pyrite was found in varying quantities in Marsa Core 1. Between the base of the core and 735cm, there is a very strong negative correlation between the amount of calcium (Ca) and the susceptibility values, which is particularly apparent for the mass specific susceptibility values χ. The Ca enrichment, which finds its peak in the calcrete deposit at 755 cm, emphasizes the overriding importance of the diamagnetic character of calcium, despite the weakness of the diamagnetic force compared

10

Although dark brown deposits had been found at the base of several cores at Malta Shipbuiliding (see above), this information had not been published.

80

noted, and from this some interesting deductions may be made.

Chapter VII BIOSTRATIGRAPHY AND PALAEOECOLOGY

7.1.2. Setting the scene 7.1. Molluscs 7.1.2.1. Data from Quaternary Deposits 7.1.1. Introduction The first description of Quaternary remains in Malta was published by A. Leith Adams in 1870, shortly to be followed by H.W. Feilden and E. Maxwell’s account on the Post-Pliocene beds of Gozo (1874). These same beds were re-investigated by J.H. Cooke in 1890, who went on to describe some Pleistocene beds in Malta in 1896.2

Any environment with calcareous sediments makes the analysis of shells a very important tool for the reconstruction of the past environment as pollen is never present in quantity in these kinds of deposits (Evans, 1975: 82), whereas mollusc shells usually are. This is even more so when sediments have been completely oxidised, as then molluscs are often the only fossils to be found (Preece, 1998: 158). Indeed, with the exception of four samples, molluscs are represented in numbers throughout the entire stratigraphical sequence of Core 1 at Marsa. Unlike palynomorphs, molluscs can often be identified down to species level and in many cases a fragment of the shell may be quite sufficient for identification. This is because some molluscs have very distinctive shell characteristics that are not shared with other species, and depending on the size of the fragment, identification can then be made at least to genus level. Juveniles, however, are often very difficult to identify, as some characteristics are only shown when the animal is adult, or at least in an advanced stage of growth (Janus, 1958: 20). Generally, it is possible to at least assign these to a specific family, but when very young, identification may be downright impossible. Due to the topography, the molluscs retrieved from Marsa Core 1 allow a general picture of the regional environment to be drawn. Additionally, the variations in the composition of the different assemblages in the samples provide a strong local signal of changing conditions.

The Quaternary deposit at Il-Ponta tal-Marfa (Marfa Point) discovered in 1935, contained not only many different land snail species, including some that are distinct from those found on the islands today, but also many species of marine shells, albeit all finely comminuted (MAR 1935-6:XXXI). Unfortunately, the results of the analyses of the molluscan remains from this deposit were never published. Until this date, few data on the Maltese malacofauna had been accumulated (see Appendix II, Table 7.1.1.), this changed with Trechmann’s 1938 paper on the Quaternary conditions in Malta, in which he also published an accurate study of the land shells occurring in local deposits (pp.1-26), among which he listed (and described) several species that he believed to be no longer present in the Maltese Islands including some new to science. From his research of various fissures, deposits and caves, Trechmann concluded that the land shell fauna is of a later date than the main mass of bone deposits, as these do not occur together as would be expected if they had been contemporary (ibid.: 16).3 Since Trechmann’s 1938 publication, Quaternary deposits containing molluscs hardly received any attention until the description in 1985 of some land snails from a Quaternary deposit at Mellieha (Thake, 1985: 93). However, the most important study of fossil shells of non-marine molluscs from a wide variety of Quaternary deposits in the Maltese Islands was made by Giusti et al. (1995). Comparing Trechmann’s 1938 taxa with their own fossil material, these authors found that only one of Trechmann’s supposedly extinct species

In this context, additional valuable information about the past environment of the Maltese Islands and anthropogenic activity through the ages can also be obtained by analysing the results and data gathered by naturalists since the 19th century through their investigations of the Maltese Quaternary deposits, and of the molluscan species that were unearthed during archaeological excavations since the early 20th century. Some limitations, however, do exist: while specimens collected from archaeological deposits can be roughly dated through their context, this is unfortunately not possible to the same extent with molluscs from Quaternary deposits, which lack archaeological remains.1 Also, usually nothing is known about the method used to collect the specimens and quantities are rarely given. At times, the lack of species smaller than 1cm appears to indicate a bias to collect larger species only. Still, although the absence of a species may be due to the collection protocol, at least the presence can be positively

2 See Giusti et al., 1995: 87-90 for a brief history of malacological research in the Maltese Islands 3 This seems to be confirmed with the 1965 discovery of a Quaternary deposits in Mriehel (Zammit Maempel, 1965: 8-10), where no shells were recorded, but plenty of bones of extinct animals. This is also the case with the Ta’ Kandja fissure (Despott, 1925-26:IX; ibid., 1926-27: XII; ibid., 1928-29: VII-VIII), the fissure at Hal Resqun (Despott, ibid., 1929-30: XIII), the Tal-Gnien fissure (Baldacchino, 1935-36: XXXI), the fossiliferous Pleistocene deposit at Ta’ Vnezja (ibid.), the ossiferous fissure at Birkirkara and the Pleistocene deposit at Hamrun (MAR 194647: IX). The collection of animal remains excavated by George Sinclair (Despott, 1923-24: X) appears to be part of the same fissure that was discovered and excavated by Baldacchino at Burmeghez in 1935, where largely the same animal remains were found in close association with Punic potsherds. Interestingly, some of these bones represent animals no longer found in Malta, like fox and small deer (Baldacchino, 1935-6: XXX).

1

Although Quaternary deposits may include remains from the Holocene, the term is hereunder opposed to ‘archaeological deposits’, because the latter may be closely dated through the material remains of past societies, while the former only contain floral and/or faunal remains.

81

Table 7.1.1. summarises the different land snail species collected from Quaternary deposits in Malta and Gozo by the various scientists. Grouping these species according to their habitats5 it emerges that ubiquitous species represent the largest group (42%), followed by steppe/open country species (35%). Molluscs that are associated with freshwater and freshwater wetlands account for a staggering 16%. It is perhaps surprising that only 4% of the species collected from Quaternary deposits are associated with leaf litter. Subterranean snails account for the remaining 3% of all species. 7.1.2.2. Data from excavations until 1971

archaeological

While the presence of marine and nonmarine molluscs in Quaternary deposits is due to natural processes, in archaeological deposits there is a strong possibility that their presence is directly or indirectly related to anthropogenic activity. There is little doubt that most marine (and brackish-water) shells found within archaeological contexts are directly related to anthropogenic activity, where they represent either food remains or ornaments. Hence, when in 1903 Themistocles Zammit was appointed Curator of Museums, the Museum Annual Reports (MAR) that were henceforth published, noted the presence of marine molluscs6 and quite often the species present were identified.7 All marine shells mentioned in the MARs are shallow water species that can still be found along Maltese shores (see Cachia et al., 1991; 1996; 2001; 2004), although the thorny oyster Spondylus gaederopus is today scarce (ibid.: 113).8 The marine shells are Table 7.1.1.: Land snails found in Quaternary deposits of the Maltese Islands. (various) thus of little value for the refers to the Quaternary deposits at Marfa, Cirkewwa, Qala (Gozo) and Dwejra (Gozo). H? palaeoenvironmental study and are = Holocene?, U = ubiquitous, FW-WL = freshwater/ wetlands, LL = leaf litter, OC = open omitted in Table 7.1.2., which summarises country/ steppe, SUB = subterranean. * = no fossil record all non-marine molluscs mentioned in the MARs between 1904 and 1971, in Evans’ was in fact extinct from the Maltese Islands.4 However, 1971 survey, and by Zammit (1930: 93-4), originating they discovered another seven extinct species among the from temple sites, tombs, caves and a fissure near a cave. 42 fossil and sub-fossil taxa that they collected. Of these, It appears that land snails were only mentioned and four were species associated with freshwater wetlands or 5 very damp habitats, indicating much wetter conditions in The percentage frequency was calculated on the basis of the past in the Maltese Islands, while the other four presence/absence data only. 6 E.g. at a cistern in Corradino prisons among land snails (unidentified), extinct species are all associated with xeric habitats (Hunt bones, masonry and pottery (MAR 1909-10:6) and at the Sta. Verna & Schembri, 1999: 65).

temple in Gozo (MAR 1911-12: 1), but without determining the species. At Skorba (Trump, 1966), only the common English name of the marine shells is given (e.g. cockle, cowrie). 7 For example at Bahrija (Evans, 1971: 107), Hagar Qim (ibid.: 94) and Ta’ Hagrat (ibid.: 35). 8 According to Caruana Gatto (in Despott, 1930:3), Spondylus gaederopus was very common until around 1870, but became very rare by 1892 to the extent that he feared it would soon be extinct.

4 Xeromunda durieui however is still common along the African shores of the western Mediterranean, in southern Italy and in Cyprus (Giusti et al., 1995: 454).

82

Table 7.1.2.: Land snails from different archaeological contexts and deposits until 1971. U = ubiquitous, FW-WL = freshwater/ wetlands, LL = leaf litter, OC = open country/ steppe, SUB = subterranean. * = edible land snails.

tombs:12 their presence among debris consisting of potsherds, implements and bones without any apparent order is likely to indicate that they formed part of the Neolithic diet and were thus disposed of with other waste. Furthermore, if these species were usually met with in Maltese caves and tombs, as Zammit says, then the indications are even stronger that apart from brackish and marine molluscs, the land snails were a staple part of the diet in the prehistoric Maltese Islands, as already suggested by the molluscs discovered at Kordin (see above).

identified when they were large enough to be easily picked out and when they were found in a context that left little doubt about the anthropogenic origin. This happened for the first time at the 1908 excavation of the eastern group of ruins on Corradino Hill (Kordin E), conducted by Thomas Ashby (Ashby et al. 1913). Here, among numerous bones of domestic animals and mixed with potsherds were twelve different marine molluscan species and the land snails Eobania vermiculata, the redbanded snail, and Cantareus apertus, the goat snail9 (MAR 1908-09: E3). A larger variety of edible land snails was recovered from the excavation of the Tarxien Temples (Zammit, 1930: 93-4), the same species were also found at the Neolithic temple at Kuncizzjoni in 1938 (MAR 1938-39: vi-ix), although here two more species are listed; both may, however, be considered too small for eating.10 More details on land snails came from the discovery and subsequent excavation of a cave used for habitation in Neolithic times in Gozo.11 Three different mollusc species are mentioned: again the red-banded snail Eobania vermiculata, the decollated shell Rumina decollata and the brackish-water tolerant bivalve the Cross-cut Carpet Shell Tapes decussatus. It is curious that Th. Zammit mentions that “the shells usually met with in our Neolithic caves and tombs are here abundantly mixed with the debris” (MAR 1913-14: 2) as neither these two land snail species, nor obviously the brackish water species occur naturally in caves and

After the initial studies of the Quaternary fossils from Ghar Dalam (see above) following the pioneering studies by Issel (1866) and Cooke (1893) the cave continued to be a favourite place for excavation. Guiseppe Despott was appointed Curator of the National History Museum of the University in 1914 (MAR 1914-15) and in 1917 he conducted excavations at Ghar Dalam. In his subsequent report he gives a detailed list of all land and marine molluscs found in the different cultural layers that he excavated (Despott, 1918: 214-221). This was followed by excavation campaigns in 1933 and 1934, led by J.G. Baldacchino, who succeeded Despott as Curator of the National History section, by J.H. Cooke in 1935 and again by Despott in 1936 (see Baldacchino’s MARs between 1933-34 and 1937-38). The methods applied at Ghar Dalam appear refined: the composition of the various layers was described and according to the pottery content broadly divided into prehistoric (Neolithic to

9

The molluscs were determined by Dr. A. Caruana Gatto. Papillifera papillaris and Chondrula pupa. 11 Ghar ta-Pergla, MAR 1913-14:2. 10

12

Rumina decollata is an omnivorous xeric species. Being a predator on other snails, it hunts at night, but also eats decomposing leaf litter.

83

Borg in-Nadur) and historic (Punic to present). Possibly because naturalists and not archaeologists were excavating the layers, many more shells were recorded than from other archaeological sites (see Table 7.1.2). While here the larger land snails again may well have formed part of the diet, the presence of molluscs smaller than 1cm in the different cultural layers could indicate that they were introduced into the cave accidentally, possibly stuck to edible plants gathered for consumption or accompanying other material.

All in all, a staggering 62% of all species collected from the archaeological sites listed in Table 7.1.2. are edible (marked with an asterix). None of the species found are associated with freshwater habitats. Both steppe/open country and ubiquitous species represent the largest groups (48% and 45% respectively), while leaf litter species amount to only 6%.

The presence of edible land snails in a funerary context is very interesting, as they appear to be part of the grave goods or food offerings, as the numerous pottery vessels found in the tomb would seem to suggest. The fact that a large percentage of the land snails recovered from the Zebbug tombs showed traces of ochre (Evans, 1971: 1669) could perhaps even indicate a ritual significance. Their presence in the tomb is even more unlikely to be accidental in the absence of smaller land snail species.13 Furthermore, at Zebbug, the tombs were sealed with a 0.25m thick layer consisting of limestone chippings (MAR 1947-48: I-II). The Neolithic tombs at Xemxija also contained large edible land snails, the same variety as found in the cave in Gozo in 1913 (Ghar ta’ Pergla near Xaghra, see Table 7.1.2.). However, it is not possible to ascertain if they were part of the burial or if they had accidentally fallen into the tombs, which were not sealed. The same accidental introduction has been suggested to account for the presence of rabbit bones discovered there (Pike, 1971: 239-41).14 Interestingly, the molluscs discovered in a 1st century AD tomb at Hal Bajjada in 1936 give a completely different picture (MAR 1936-37: XXIII) – only one out of the ten different species of land snails discovered here is large enough to be considered as a food item. The overall assemblage, however, would suggest that they all were part of the top soil assemblage from the surroundings, which was used to cover the tomb.

The use of molluscs in environmental reconstructions only dates back to the last century, after Charles Elton in 1927 redefined ecology as ‘the study of animals (and plants) in relation to habit and habitat’ and thus in fact launched a new branch of biology (Evans & O’Connor, 1999: 1, 3). Before this time, the interest in molluscs was more on a taxonomic level, where researchers were mainly preoccupied with the discovery and description of species. In Malta, it is only of recent date that the value of biological components in archaeological sediments has been recognised. Environmental samples, with the intent to reconstruct the environment, were taken for the first time at the excavation of the prehistoric Xaghra Circle in Gozo, although these results still await publication (Schembri & Hunt, forthcoming). The techniques were refined and applied on a much wider scale for the subsequent excavation by the University of Malta at TasSilg, where over 120 bulk samples were taken and analysed for their artefactual, geological and biological content (see Schembri et al., 2000 for first results, and Fenech, 2001a for a case study at Tas-Silg), apart from the hand-picked molluscs that were collected from all excavated layers.16 Molluscs from the Victoria Cave in Victoria, Gozo17 (Figure 1.5.) and pollen analysis from a tufa deposit at Fiddien Valley also provide an important insight into the past environment (Hunt, 1997). An analysis of the molluscan contents from a Punic tomb at Xemxija (Fenech, 2001b) adds to the general picture (Table 7.1.3). From the percentage distribution analysis according to their broad habitat preferences, it emerges that again the ubiquitous and steppe/open country species are the most numerous 37% and 35% respectively). Fresh water species are also present, although they represent the smallest group (2%), followed by subterranean species (4%). Leaf litter species account for 22% of all species taking all sites collectively.

7.1.2.3.: Quantitative data from samples in archaeological contexts.

A fissure discovered and excavated at Burmeghez in 1935 revealed cultural material from the Punic Period mixed with many animal bones15 and land snails (MAR, 1935-36: XXIX-XXX). Whether the edible red-banded snail Eobania vermiculata, the goat snail Cantareus apertus and the garden snail Cantareus aspersus were also food remains or other waste is, retrospectively, impossible to ascertain as the ubiquitous small Chondrula pupa and the clausiliid Papillifera papillaris are also listed.

environmental

7.1.3. The environmental background The general picture that emerges from comparing the figures and Tables 7.1.1., 7.1.2. and 7.1.3. is very interesting. All in all, ubiquitous species represent the largest group found, which is no surprise as these have the least demanding habitat requirements (Schembri et al., 2000: 103). Steppe and open country species are,

13 The presence of smaller land snails would have been noted by J.G. Baldacchino, who had held the post of Curator of the Natural History section before being appointed Director of the Museum later that year. Also, since this was the first time a Neolithic burial was discovered, the excavation was “conducted with meticulous care” (Baldacchino, MAR 1947-48: I-II) 14 After the last Ice Age, the European wild rabbits were confined to the Iberian peninsula. They may have been dispersed already with the Phoenicians, but the Romans are known to have introduced the rabbit to many parts of the Roman Empire. It is thus unlikely that the rabbit remains found at Xemxija predate the Punic/Roman era. 15 The bones belonged to a small species of horse or ass, a small deer, sheep or goat, pig, small fox and tortoise.

16

Although there is a marked bias to collect large snails and miss out on many small species, they still provide important indications. 17 Dated to the Punic and Roman Period by their direct association with pottery inside the cave. The tufa deposit has unfortunately not been carbon dated.

84

assemblages permit a clear picture of the precise local environment to be drawn (Preece, 1998: 158). The coring site at Marsa, on the other hand, is within a high-energy depositional environment. Shells belonging to the following broad habitat categories have been found in Core 1: land, fresh water, brackish water and marine. 7.1.4. Preservation of shells from Marsa Core 1 Due to the high calcareous content of local soils and their alkaline nature, the preservation of the shells is generally good, but only in the sense that they have not disintegrated beyond recognition. However, the vast majority of shells sustained some degree of damage or erosion. 7.1.4.1. Fracturing Fracturing of the shell was the most common damage. This is due to different factors: land snails have a comparatively thinner shell than marine shells and thus generally fragment more readily through passive transportation, re-deposition, burial and the resulting postdepositional compression. Damage occurred usually at the most fragile parts of the shells, i.e. at the aperture and/or at the tip (apex), but completely fractured shells were also very common. The few land snail specimens that escaped fracturing were those of very small size, either juveniles or adults belonging to small species. In some cases the land snails were so severely fractured that identification was impossible. Brackish-water and marine gastropods were also often fragmented, but when transported their thicker shell prevented severe damage. For such species, shells of up to 4cm size were often found intact. This was not the case for bivalves, although many species live within the sediment and are thus not subjected to transportation damage if they die and remain in situ: only four complete specimens (articulated) were found, but none of which was larger than 4mm. Other bivalves often showed signs of predation by such epibenthic predators as crabs and fish that crush the shells in order to eat the animal inside (Barnes, 1994: 34). Although severely fragmented, identification was in most cases possible at least to family level.

Table 7.1.3.: Land snails from environmental samples of archaeological deposits. U = ubiquitous, FW-WL = freshwater/ wetlands, LL = leaf litter, OC = open country/ steppe, SUB = subterranean

however, nearly of equal importance to the ubiquitous species in all three tables and figures. Although there is no absolute date for any of the Quaternary deposits, it is reasonable to assume that these deposits probably predate the arrival of the first settlers on the Maltese Islands and thus would reflect an ‘environment undisturbed by humans’, which appears, from the data, to have been predominantly open country/steppic. Compared to the data from Tables 7.1.2. and 7.1.3., the largest species diversity was shown by the Quaternary deposits. Interestingly, apart from one mention of Pleurodiscus balmei by Adams, leaf litter species are conspicuous by their absence, which could substantiate the suggestion of a generally open country/steppe environment. Freshwater and freshwater wetland species appear to be much more common and diverse in the Quaternary deposits than in the archaeological deposits, although this may be due to the different nature of the various sites. Only in some low-energy depositional environments is it possible to recover assemblages that are autochthonous, i.e. the assemblage is uncontaminated with shells from different deposits, spatially and temporally. These 85

7.1.4.2. Blackening

spratti. When there was doubt, these were also added to the Heliciidae.

A small percentage of the brackish-water and marine shells showed grey to dark grey mottling, while a few had turned completely black. The blackening is a common reaction of the shell in anoxic and reducing conditions and is due to the deposition of metal sulphides. It can occur during the mollusc’s lifetime but also after the death (Schembri, personal communication: 2005). The blackening did not affect the identification.

Hydrobia spp. Three species belonging to the family of hydrobiids occur in the Maltese Islands, Hydrobia acuta, Hydrobia ventrosa and Heleobia stagnorum. These can occur together as they have very similar habitat requirements. Since easy differentiation between the three can only be done with the live male animal inside the shell (Giusti et al. 1995: 116-125), identification was only made to genus level.

7.1.4.3. Erosion and weathering of the shells Cerithium spp. It is not only very difficult to distinguish a juvenile shell of Cerithium vulgatum from an adult shell of Cerithium rupestre but also when both are juvenile. For this reason, shells belonging to either species were grouped under Cerithium spp.

Several brackish-water and marine shells showed evidence of having been rolled by water action over rough ground or abrasive sediment and as a result the shell became at times heavily eroded. Some of these shells were also weathered by the sun and had lost their colouring. These damaged shells are likely to be allochthonous to the shell assemblages as their life provenance would have been some indeterminable distance away from the coring site. Furthermore, their presence is not necessarily contemporary with the live assemblage. As such, they were still sorted and analysed at least to genus level, but are not presented in the data tables.

Cerastoderma spp. Two species of the cockle exist in the Maltese Islands: the edible cockle Cerastoderma edule and the lagoon cockle Cerastoderma glaucum. While the latter can occur in very low salinities (down to 4 on the practical salinity scale), C. edule needs a minimum salinity of 15. C. edule replaces C. glaucum in estuaries in which it can occur (Barnes, 1994: 148-9). Although both species usually have distinguishing characteristics, certain identification was often impossible from the fragmentary remains. Furthermore, several shells were found that had characteristics of both C. edule (short radiating ridges on the interior) and C. glaucum (horizontal ridges between the radiating ridges). As both species occur in brackish water habitats, they were recorded as Cerastoderma spp.

7.1.5. Notes on identification In addition to the difficulties caused by poor preservation, several juvenile species and the lack of the live animal in the shell also caused considerable problems concerning the identification of some taxa. Ferussaciidae

Bivalvia Two different species belonging to this family could be identified in Core 1: Ceciliodes acicula and Hohenwartiana hohenwarti. Differentiation between the two could only be made when the aperture of the shell was intact. Otherwise, the broken individuals were grouped under ‘Ferussaciidae’.

Although bivalves occur naturally in pairs, they usually occurred singly or in uneven pairs in the samples of Marsa Core 1. Each valve was counted as one individual (after Ridout-Sharpe, 1998: 338). 7.1.6. Nature of the molluscan assemblages

Clausiliidae Although the area of the coring site shows characteristics of a low-energy environment, where, if changes happen, they happen slowly, the environment is also subjected to sporadic high-energy events. However, with high-energy events like heavy rainfalls and storm floods, molluscs from different habitats are easily carried with sediments and other debris into or onto the plain where they are redeposited. These allochthonous molluscs are then mixed with the autochthonous assemblage but it is at times very difficult to establish which molluscs have been washed in. Furthermore, the rising of the sea level at the end of the last glaciation submerged a former land environment and changed it to a brackish-water/marine one. This means also that a land environment, where land snails died in situ would have changed to a brackish-water or marine environment with subsequent replacement of the molluscan fauna, through simple rising of the sea level.

The characteristic ribbing of the genus Muticaria could be identified even from tiny fragments. On the other hand, it was often impossible to distinguish between juvenile Papillifera papillaris and Muticaria sp. Doubtful cases were listed under Clausiliidae. Heliciidae When adult, helicid shells were usually heavily fragmented to such a stage that it was impossible to identify the species. Here, the different apexes were counted and all grouped under ‘Heliciidae’. Included here are Eobania vermiculata, Cernuella spp., Marmorana melitensis and Cantareus spp.. Furthermore, in their juvenile stage it is often impossible to differentiate between the shells of Theba pisana and Trochoidea 86

Both shell groups would be autochthonous, although not contemporary.

which corresponds closely to the values formerly expressed in parts per thousand of salt (McLusky & Elliot, 2004: 3). The Venice System (1959) classifies water bodies on the basis of their salinity in the following way: 18)

(salinity > 30)

0 10

20

30

40

50 60 0

20 30 40 50 60 70 80 90

0

20 30 40 50 60 70 80 90

0

1

2

3

4

total number of shells 5 0

1000

2000

0

1

2

3

depth in meters

4

5

6

7

8

9

10

11 PANGAEA/PanPlot

max.: 11.15

Figure 4 - land-halines species.txt - 05.04.2005 17:16 h

Figure 7.1.1.: Variations of habitat specific assemblages throughout Marsa Core 1, expressed as percentages of the total number of shells.

87

adopted here groups the species according to landscape/ environment into the following habitats mentioned in Giusti et al. (1995): open country (steppe and garigue, including xeric habitats that do not receive substantial water during the dry season and only have little shade), leaf litter (dark, shady habitats with some moisture), ubiquitous (found in a wide variety of both natural and anthropogenic environments), subterranean (burrowed in soil or under deeply embedded stones) and freshwater habitats.18 Some land snails live on vegetation (e.g. the goat snail Cantareus apertus, see Giusti et al.: 1995: 488490), but this is also the case with many ubiquitous snails. Therefore, this category was not used here. When grouping all the land snail species found in Marsa Core 1 according to these broad habitat categories it appears that compared to the previous figures from the Quaternary deposits and archaeological contexts, most habitat categories show a larger percentage of different species, thus minimising the ubiquitous species group (24%). The freshwater species represent the largest group (28%), while open country and leaf litter species both correspond to 20% of all species. The smallest group here are the subterranean species (8%). The picture changes considerably when listing the species by numbers present and habitat categories (standardised on 150g/sample). The vast majority of land snails found belong to

Table 7.1.1.: Molluscan species according to their salinity ranges.

euryhaline and polyhaline species have been added together to form the brackish-water group. From here it becomes very clear that the brackish water group is by far the largest, while the presence of marine species is the exception. The total number of shells also shows the variations in the abundance of molluscan remains throughout the core (Figure 7.1.2.).

18

Another method of grouping is to sort the species according to their temperature/moisture preferences as in Schembri et al. (2000:103-4) and followed by Fenech (2001b: 20-22): ubiquitous (as above), xeric (habitats that do not receive a substantial amount of water during the dry season), mesic (habitats that do not remain dry for long periods even during the dry season), subterranean and freshwater. The difference, however, between the two ways of grouping is minor as country/steppe broadly corresponds to xeric habitats as do mesic to leaf litter.

7.1.7.2. Land snails according to their habitat requirements There are several ways of grouping land snails according to their broad habitat preferences. The system that has been 88

land snails

freshwater molluscs

(%) 0

(%)

20 30 40 50 60 70 80 90

0 10

20

30

euryhaline species

polyhaline species

(%) 40

50 60 0

marine molluscs

(%)

20 30 40 50 60 70 80 90

0

total number of shells

(%)

20 30 40 50 60 70 80 90

0

1

2

3

4

5 0

1000

2000

0

1

2

7.1.7.3. Brackishwater and marine molluscs according to benthic habitats

3

4

depth in meters

half of the core, while molluscs associated with leaf litter are generally very scarce and occur only sporadically in very low numbers.

5

6

7

8

9

10

11 PANGAEA/PanPlot

max.: 11.15

Figure 7.1.4 - land-halines species.txt - 22.05.2006 09:52 h

Due to biological activity, the mud deposits of Marsa Core 1 are likely to be intensely bioturbated (see Chapter VI). The ground disturbing activity is not confined to molluscs only;

among other agents, soft bottoms are also inhabited by polychaetes and various crustaceans, which build their burrows in the mud, and by small crabs, fish, echinoderms and other molluscs that live off the bivalves, all manner of buried fauna, carrion and detritus. This is important to note as the stratigraphy may thus be disturbed to a certain degree and communities within one sample are unlikely to be homogenously contemporary. Furthermore, datable plant material might be residual or intrusive.

Figure 7.1.2.: Variations of molluscan assemblages throughout Marsa Core 1 according to their broad habitat category, with euryhaline and polyhaline species grouped together as brackish water.

ubiquitous species (69%), while open country species represent 11%, and leaf litter only 1% of all land snail species found. Freshwater molluscs are the second largest group, with 14% of all non-marine molluscs, while subterranean species account for 5%.. Figure 7.1.3. shows the variations in the land snail distribution according to their broad habitat requirements throughout Marsa Core 1, standardised per 150g/sample. Freshwater molluscs predominantly appear in the upper

The brackish-water and marine molluscs can be divided into three groups: epifaunal, epi/infaunal and infaunal, according to their habitats. Epifaunal ubiquitous open country leaf litter fresh water Total number of shells molluscs (e.g. gastropods and oysters) live either directly on the sea floor or are attached to some secondary substratum like rocks or stones or other shells, while epi/infaunal molluscs live on the sea floor, but also bury shallowly in the soft bottoms (e.g. the cockle Cerastoderma glaucum). Soft bottom mud offers a habitat also to detritus feeding bivalves where they can burrow deeply. Figure 7.1.4. shows the presence and distribution of Figure 7.1.3.: Land snail distribution throughout Marsa Core 1 according to broad habitat category, expressed 0 10

20

30

40

50

60 70 0 2 3 4 5 6 7 8 9 11 13 15 17 19 0

1

2

3

4 0

10

20

30

40

50 0 20 40 60 80 100 130 160 200

0

1

2

3

depth (m)

4

5

6

7

8

9

10

11

PANGAEA/PanPlot

max.: 11.15 m

Figure 7.1.6..txt - 22.05.2006 10:27 h

in numbers per 150g sample.

89

(Giusti et al., 1995: 304, 308). With regards to the terrestrial samples of Marsa Core 1, it is difficult to ascertain when they should be treated as intrusive. On the other hand, in assemblages where they are clearly allochtonous, they may provide valuable evidence of erosion. Figure 7.1.5. shows the presence and distribution of these Ferussaciids throughout the core.

Percentage distribution epifaunal

epi/infaunal

(%) 0 10 20

30 40 50

infaunal

(%) 60 70

80 90

0 10 20

30 40 50

total sum

(%) 60 70

80 90

0 10 20

30 40 50

60 70

80 90

0 200 400 600 800

1100 1400 1700

0

1

2

3

depth (m)

4

5

6

7

8

9

10

Shallow and deep burrowers (non-marine

11 PANGAEA/PanPlot

max.: 11.15 m

in-epifaunal amended.txt - 19.05.2006 11:45 h

and marine) occur throughout Marsa Core 1, and while their allochthonous presence is likely to be indicative of bioturbation prior to the erosive events that removed and redeposited them together with the sediments at the coring location of Marsa Core 1, their autochthonous presence would be indicative of in situ bioturbation.

Figure 7.1.4.: Distribution of epifaunal, epi/infaunl and infaunal marine molluscs in Marsa Core 1, standardised per 150g sample and expressed as percentage values.

epifaunal, epi/infaunal and infaunal molluscs in Marsa Core 1. 7.1.7.4. Burrowing land snails Terrestrial soils are also subject to biological activity. Both Ceciliodes acicula and Hohenwartiana hohenwarti are burrowers. While Ceciliodes acicula commonly occurs in light friable soils at depths up to 1.5m (Preece, 1998: 202), Hohenwartiana hohenwarti locally prefers to burrow deep under stones in the Blue Clay deposits

7.1.7.5. Number of species

Subterranean species 0

1

2

3

4

5

0

1

2

3

depth in meters

4

5

6

7

8

9

10

11

The number of species found in each sample of Core 1, Marsa, varies greatly throughout the length of the core, between 1 species and a maximum of 30 different species. Since these species belong to up to five different broad habitat categories per sample, which indicates the allochthonous nature of several of these molluscs within the samples, a calculation of dominance, evenness or similarity as usually done in ecological studies, was not applied here. Instead, a simple comparison of the number of species per broad habitat and sample reveals that the polyhaline group has by 6

7

8

PANGAEA/PanPlot

max.: 11.15

Ceciliodes panpl.txt - 22.05.2006 13:21 h

Figure 7.1.5.: Distribution of burrowing land snails belonging to the family of Ferussaciidae throughout Marsa Core 1, standardised per 150g sample. Their presence need not necessarily always imply bioturbation as they may have been part of a hillwash sediment deposited at the Marsa Sports Ground.

90

far the highest species diversity, while the marine species are the

The community structure of the assemblages may be land snails freshwater molluscs euryhaline species polyhaline species marine species total species total number of shells studied by non-metric multi-dimensional scaling (MDS), based on a similarity matrix constructed using the Bray-Curtis similarity index using the PRIMER v5 software. Figure 7.1.7. shows similarities and variations of the molluscan assemblages of all samples from Marsa Core 1. The shells were grouped by broad habitat categories (land, fresh water, brackish water and marine), standardised per 150g/sample. The resultant plot shows a distinct group at the top, which contains the Figure 7.1.6.:Species richness and abundance in Marsa Core 1. The largest number of species occur in the predominantly brackish polyhaline group, while the marine species are comparatively few. water assemblages. The group at the bottom centre least diverse. Compared with the total abundance of of the graph are assemblages without any brackish water shells, the euryhaline group is, despite the low number of or marine molluscs, while the stretched grouping above different species, considerably more dominant in numbers contains predominantly various amounts of land snails. than the diverse polyhaline group. Figure 7.1.6. also The complexity and variations of the assemblages demonstrates two things: the number of species is not throughout Marsa Core 1 is emphasized by the numerous related to the number of shells found; and that land snails outliers and a general paucity of recognisable sequences and freshwater shells appear prominently in a large in the sample compositions. A second analysis was run number of the various samples, highlighting the with a simplified matrix based on presence/absence of allochthonous nature of the assemblages. molluscs belonging to broad habitat categories (Figure 7.1.8.), which resulted in 12 groups. Of these, two were 7.1.7.6. Statistical analyses clear outliers as they contained only one sample each. 0

2 3 4 5 6 7 89 0

1

2

3

4

5 0

1

2

3

4

5 0 3 5 7 9 11 14 17

0

1

2

3

4 0

10

20

30

40 0

1000

2000

0

1

2

3

depth (m)

4

5

6

7

8

9

10

11

PANGAEA/PanPlot

max.: 11.15 m

species richness core 1 ready.txt - 22.05.2006 13:43 h

Figure 7.1.7.: Non-netric multidimensional scaling ordination of the molluscan assemblage from all samples from Marsa Core 1, standardised per 150g/sample, based on a similarity matrix constructed using the Bray-Curtis similarity index. The plot shows a distinct group at the top, which are predominantly brackish water or marine molluscs, the less dense assemblages towards the right contains predominatly land snails in differing amounts. The complexity and variations of the assemblages throughout Marsa Core 1 is emphasized by the paucity of distinct groups in the ordination diagram.

91

Figure 7.1.8.: Non-metric mulitdimensional scaling ordination of the molluscan assemblages from Marsa Core 1 based on presence/ absence of broad habitat types (land, fresh water, brackish water, marine), in the samples. The simplified grouping resulted in 12 sub-groups, of which 1/196 and 1/024 are the clearest outliers. 1/196 contained only 1 marine shell, while 0/24 contained land, freshwater and marine shells, but no brackish water shells. The two closely aligned groups at the centre of the plot are again of the predominantly brackish water assemblages, where the top group contains no freshwater molluscs and the group below no marine molluscs

1/196 consisted of only a marine shell, while 1/024 contained land, fresh water and marine shells, but no brackish water molluscs. The most common groupings, shown at the centre of the plot, consisted of predominantly brackish water assemblages that had either no marine or no fresh water molluscs.

transportation of considerable amounts of sediment through flooding. The lack of fragmentation in some shells may be indicative of no post mortem transportation and hence such shell may be a component of the autochthonous assemblage. The total number of land snails varies greatly throughout the core, but is generally highest in the upper 340cm of the core.

7.1.8. Analysis of faunal change throughout the core The different land snail species shows no particularly great diversity, but this is likely to be due to the sample sizes. With an increase in the overall number of land shells in the upper 340cm of the core, the number of species also increases (see Figure 7.1.6.).

For the subsequent analyses of the various graphs, the uppermost 25cm of the core will not be considered as these sediments were disturbed through the previous excavation and ground levelling activity at the coring site. Grouping the molluscs into four broad habitat categories as shown on Figure 7.1.2., gives a very complex picture (also emphasized by Figure 7.1.7.). Most habitat categories are present throughout the core, as if the environment changed several times between land and marine. This appears to emphasize the quasi permanent admixture of allochthonous and autochthonous molluscs throughout Marsa Core 1. Particularly in the lowermost 4 m of sediments of Marsa Core 1, from 1120cm to around 730cm, where the colour of the sediments would indicate that they are mainly terrestrial, the presence of brackish water and marine molluscs may be the result of minor wash-over events. The few freshwater molluscs that occur in this part of the core were all embedded within grey coloured sediments, possibly as a result of major wash-over events.

Of all the different shell groups, the greatest diversity of molluscan species is found in the polyhaline molluscs, particularly from 3.4m down as can be seen in Figure 7.1.6. Marine molluscs do not exceed three different species at any point, which may perhaps be indicative of the predominantly brackish water environment. The marine molluscs may perhaps be allochthonous, although it is more likely that fluctuating salinities would have enabled the few marine species to find their way actively into the deposits. On the other hand, the freshwater species would have ended up in the deposit through rainfall events that would have created currents strong enough to dislodge the fresh water molluscs and wash them into the lagoon. Figure 7.1.4. shows that bioturbation of the sediments by burrowing marine species appears to be particularly pronounced between 700cm and 290cm as indicated by the comparatively high total sum of epi/infaunal and infaunal molluscan species. In this part of the core, several disarticulated dactyli of crabs were also found, which further points to intense bioturbation of the sediments. The presence of epi/infaunal and infaunal marine molluscs even among the non-marine sediments in the lower part of the core suggests a reworking of these sediments even without the activity of bioturbating molluscs.

Between 700cm and 340cm, there is an overwhelming dominance of brackish water molluscs, with a few marine shells and very few land and freshwater shells. The total shell sum is highest in this part of the core, which consisted here of grey coloured silts (compare Appendix I), indicating a calm brackish water environment into which land and freshwater molluscs were only occasionally washed. The situation appears to change radically at 335cm, when the percentage distribution of the molluscs from different habitats is altered. An increase in land and freshwater molluscs is coupled with a progressive decrease in brackish water and marine molluscs. The drastic decrease in the total shell sum may be indicative of increased sedimentation rates, possibly due to the silting up of the lagoon. This is further indicated by a progressive increase in land shells. The variations in the number of freshwater molluscs may, perhaps, be indicative of the magnitude of the flooding events, but in the upper part of the core, particularly from 1m to the top, it may also suggest a disappearing fresh water body.

In conclusion, it emerges from the above observations that the vast majority of the molluscan assemblages within Marsa Core 1 contain a varying degree of allochthonous shell material. From around 730cm upwards it is fairly straightforward to establish which shells are part of the original assemblage and which are not as these sediments are marine. The lowermost 4 m of the core, however, appear considerably more complex. While here the assemblages in the grey marine sediments (at 800cm – 805cm and at 830cm – 880cm) are comparable to those between 730cm and 340cm, the same cannot be said for the various red coloured sediments in the lower part of the core, which also contain similar assemblages, albeit in much smaller numbers. It appears very likely that several sediments may have been reworked a number of times in the past. This is also indicated by the radiocarbon dates in the lower part of the core that produced the zigzagged timeline (see above, Chapter V and Figure 5.2.). Consequently, any components within the sediments of

Figure 7.1.3. shows the percentage distribution of the land and freshwater molluscs according to their broad habitat requirements. Ubiquitous shells form the largest group, but open country species are also present from the lowermost part of the core onwards, while molluscs associated with leaf litter form the smallest group and only occur very sporadically. The presence of the subterranean species throughout the core may in many cases be allochthonous and hence point to the 92

the lower part of the core may have been reworked and is possibly not contemporary with its last burial that resulted in the final stratigraphy of Marsa Core 1.

its habitat, but the very few fossil specimens were found at Wied tal-Bahrija, which indicates that well-vegetated wet banks could have provided a suitable habitat. According to Giusti et al. (ibid.) only a single recent specimen has been found so far, in the channel of a permanent spring in Gozo. Its possible extinction from Malta, in conjunction with the disappearance of, among other, Planorbis planorbis and Bulinus cf. truncatus has been hypothesized by Giusti et al. to be probably due to a hot dry post-glacial period which would have dried up the stream at Wied il-Bahrija. This would have seriously disrupted the ecology, but with a change in climate the stream would have been re-colonised by freshwater molluscs, most likely through passive dispersal (ibid.: 162, 180, 200). While this hypothesis may be true for Wied il-Bahrija, evidence from Marsa Core 1 would suggest that Carychium cf. schlickumi and Bulinus cf. truncatus were present in Malta at least until the Late Bronze Age. The reason for their disappearance may be climatic, but it is also possible that an anthropogenic influence may have negatively affected the freshwater ecology at Marsa.

7.1.9. Notes on selected species The vast majority of molluscs recovered from Marsa Core 1 still form part of the present day Maltese fauna. However, some freshwater species are very localised today due to the overall dearth of freshwater on the islands, while other freshwater species have become extinct. These are discussed below. Planorbis planorbis This species inhabits muddy stagnant or slow-moving freshwaters that are rich in vegetation (Giusti et al., 1995:180, Janus, 1958:48). It has a Holopalaearctic distribution and has been recovered from several Quaternary and archaeological deposits in the Maltese Islands (see above), but is today locally extinct. In Core 1, Marsa, Planorbis planorbis appears frequently among the other freshwater species, but is often also the only freshwater species found. It appears for the last time in Sample 44 and by Sample 33 it is replaced by Planorbis moquini, a member of the same genus and which has the same biogeographical distribution and habitat requirements, except that it prefers sandy substrata in more or less permanent freshwater bodies (Giusti et al., 1995: 184). This species still occurs in the Maltese Islands, although due to the rarity of its habitat, Planorbis moquini is very scarce.

7.2. Vertebrates 7.2.1. Introduction Remains of vertebrates were very scarce throughout the core. While some of the bones were in a good state of preservation, others were reduced to tiny fragments, which at times showed also signs of weathering. Despite the good state of preservation, it was impossible to determine the species or even genus for a tibia or a femur other than ‘vertebrate’ or ‘mammal’. Teeth may be more diagnostic than bones: several molars of a rodent were also discovered, in a good state of preservation, as was a fragmented mandible of a lizard with some teeth still in situ.

Bulinus cf. truncatus This species lives in a wide variety of freshwater bodies, which may also be seasonal. The few fossil specimens discovered by Giusti et al. (1995) came from the banks of a permanent freshwater pool in Gozo (Ta’ Sarraflu) and the banks of the perennial watercourse at Wied il-Bahrija. Its geographical distribution is confined to the Mediterranean, the Middle East and large parts of the African continent, but it is now no longer found in the Maltese Islands (ibid.: 1995: 198). Eight individuals were found in Marsa Core 1, all of which were juveniles. Bulinus cf. truncatus appears as low down as Sample 145 and did not occur beyond Sample 95, although other freshwater species were found nearly all the way up in the core. Since Sample 90 has been dated to 800-500 cal. BC (94.2% probability), the last appearance of Bulinus cf. truncatus in the core may possibly correspond to the Late Bronze Age/Early Phoenician period, as the sedimentation rate in this part of the core appears to be very low (compare Figure 5.2.).

7.2.2. Identification and results The material was identified by Prof. Wighart von Koenigswald at the Palaeontological Institute in Bonn, while the lizard mandible was identified by Prof. Wolfgang Böhme of Museum Koenig in Bonn. The following remains of vertebrates were found: The bones found in the core all belonged to small animals (see Table 7.2.1. below). None of the longbones or longbone fragments were larger than 4mm, but the horn fragment was measured 1cm x 0.6 cm. The bad state of preservation of some of them may be due to predators or scavenging activity. The weathered fragment of a horn belonged to a mammal, but the tiny size of the fragment makes it impossible to identify the species. As it is impossible to identify the various bones more precisely, their remains do not add to the environmental history at Marsa.

Carychium cf. schlickumi One specimen of this species has been found in Marsa Core 1 (Sample 96). Presently, this species has a Transionian distribution, although it has been found in Pliocene deposits in the Rhineland and the Caucasus, and in Plio-Pleistocene deposits in Dijon in Central France (Giusti et al., 1995: 162). Generally, Carychium cf. schlickumi prefers a freshwater wetland habitat in warm regions (Pfleger, 1984: 64). In Malta, nothing is known of

Identified species: A mandible of the Maltese wall lizard Podarcis filfolensis was discovered at 320cm in an early Byzantine context. 93

The allochthonous presence of this very common lizard in a brackish water sediment adds, however, little knowledge to the present study.

7.3. PALAEOBOTANY

The most interesting find was the presence of two wellpreserved molars from the lower mandible of the extinct vole Pitymys melitensis. Several specimens of this vole had been discovered by Bate in 1920 at Ghar Dalam cave and again by Storch during his 1969 excavation at the same cave. Although some remains of Pitymys melitensis had been found in the cultural layer at Ghar Dalam cave, their appearance and the scattered occurrences of the teeth made it clear that they were intrusive (Malec & Storch, 1970: 76; Storch, 1970: 136) and that they belonged to the Red Deer layer, dated to the Late Pleistocene (Zammit Maempel, 1989:36-7; Hunt & Schembri, 1999: 51,53, Storch, 2006: personal communication). The extinction of this vole is probably due to the disappearance of deep soils and/or increased seasonal droughts that lead to severe desiccation of the soils (Malec & Storch, 1970: 78). The presence of the molars of Pitymys melitensis within the dark red soils is a further indication of the old age of the sediments towards the base of the column, as already indicated by the radiocarbon dates (see above). Whether the two molars belonged to the same or to two different individuals cannot be ascertained at this stage. If they are from one individual only, the difference of 25cm between the findspots within the column may indicate a reworking of the sediment in the lower part of the core. This may be due to bioturbation, or it could indicate a secondary deposition of sediment through mass transport, which, however, would make the occurrence of 2 molars of the same vole at that vertical distance a chance occurrence.

In order to assess any anthropogenic influence on the environment of the Maltese Islands, it is necessary to reconstruct the past environment from the available data. The environment, however, has changed considerably since the appearance of the first plant on the Maltese Islands, mainly as a result of regressing and transgressing sea-levels.

7.3.1. Introduction

It appears that the Maltese Islands would have received their initial influx of biota during the Messinian via broad land-bridges from both southern Europe and northern Africa, although the North African influence may be, if at all, minor (Hunt & Schembri, 1999: 68-9), as several studies show clear affinities of the Maltese Quaternary biotas with Sicily (e.g. Francini Corti & Laza, 1972; Lanfranco, 1984; Schembri, 1992) Particularly during the Pleistocene regressions, when the Maltese Islands were either connected via a land-bridge with Sicily or else only separated by a narrow channel, biotic immigration waves occurred, which seem to have replaced the previous biota, as evidenced by the faunal remains at Ghar Dalam cave (Schembri, 1992: 70). Similar successions are likely to also have occurred with the flora, but little is known about the Pleistocene vegetation of the Maltese Islands. Pollen analysis from a possibly mid-Pleistocene tufa deposit at Fiddien Valley shows evidence of lacustrine and lake-marginal vegetation, surrounded by scrub and some open ground taxa, typical for an interglacial period (Hunt, 1997: 1023). Other undated Pleistocene deposits revealed the presence of leaves from the laurel Laurus nobilis and the Aleppo pine Pinus halepensis (Zammit Maempel, 1977; 1982).

7.2.3. Conclusion The bone remains of the vole Pitymys melitensis reinforce the presence of Pleistocene elements within the sediments at the base of the core and the occurrence of mass transports. Apart from this, the paucity of identifiable vertebrate remains adds little to the reconstruction of the palaeoenvironment.

Studies about what exactly the Maltese Islands looked like shortly before the first permanent settlers introduced agriculture have, so far, not been based on environmental evidence from Malta. More is known about the natural early Holocene vegetation since the Neolithic. Excavations at Skorba revealed the Sample Depth (cm) Specimen Species Period presence of the Judas tree (Cercis siliquastrum), Hawthorn (Crataegus 58 320 mandible Podarcis filfolensis early Byzantine sp.) and Ash (Fraxinus sp.), which 59 325 ?metatarsal, femur vertebrate early Byzantine were used for firewood (Metcalfe, 73 425 hair Equus sp. Roman 90 510 longbone fragment vertebrate Phoenician/Punic 1966). Pollen analysis of a Bronze Age 96 540 longbone fragment vertebrate Late Bronze Age pit at Tal-Mejtin near Luqa included a few pine and olive pollen grains among 98 550 longbone fragment vertebrate Bronze Age predominantly herbaceous families 166 960 femur vertebrate > Neolithic (Godwin, 1961). That agriculture was 183 1045 molar Pitymys melitensis > Neolithic practised since the Neolithic (see 187 1065 femur vertebrate > Neolithic above, Chapter III) is evidenced by 188 1070 molar Pitymys melitensis > Neolithic various studies of macrobotanic 190 1080 longbone fragment vertebrate > Neolithic remains in archaeological deposits 190 1080 horn mammal > Neolithic (Helbaek, 1966; Renfrew, 1972; Fenech, 2001a) and pollen analyses Table 7.2.1: List of vertebrates remains found in Marsa Core 1. (Godwin, 1961; Hunt, 2000).

94

The present flora of the Maltese Islands is popularly regarded as being impoverished, but considering the relatively small land area, the limited number of habitat types and intense human activity, the islands harbour a very diverse array of plants (Schembri, 1993: 32). Of the over 1700 (estimated) known plants in the Maltese Islands, only a fraction is reflected by a study of the palaeobotany.

eliminated; this was determined when the preparation stopped bubbling. The remaining sediment on the mesh was washed with a fine spray. Occasionally, this material was ‘swirled’ on the clock-glass to continue the process of removing heavier particles. When eventually the filtrate was clear, it was stained with safranin and pipetted into a labelled 10cm³ sample tube and stored in a refrigerator overnight to settle. The next day, the supernatant water was pipetted off, and a few drops of the sample at the bottom of the tube were placed on a labelled microscope slide on a hotplate set at 100° C. After adding two drops of Aquamount the product was mixed together with the corner of a cover slip prior to the cover slip being mounted. These slides were left to set for at least 24 hours before detailed examination was made.

7.3.2. Pollen Analysis Very few pollen studies have so far been undertaken in the Maltese Islands (see above). Pollen analyses are very useful tools in reconstructing past environments and have also been applied successfully in numerous studies in the Mediterranean (e.g. Carrión & Van Geel, 1999; Ramrath et al., 2000; Carrión et al., 2003; Sadori et al., 2004).

As the total pollen count was disappointingly low, Carroll decided to re-treat each pollen sample by decanting the stored concentrate into a 100cm³ beaker, making-up to 50cm³ with distilled water, adding 5g of sodium pyrophosphate and boiling for 30 minutes. Subsequent sieving, swirling, pipetting and staining were carried out as outlined above before storing in 10cm³ tubes overnight in a refrigerator. Slides were again prepared and, upon reexamination, the results showed a marked increase in the counts of pollen, spores and palynofacies. This is attributed to a further breakdown of clays by sodium pyrophosphate.

In a Maltese context, however, there are several disadvantages inherent with pollen preservation. Pollen decays in alkaline, open textured sediments, as the high pH value of the deposits favour attacks by microorganisms (Evans & O’Connor, 1999: 134). Furthermore, in deposits that were biologically active, this would have encouraged direct chemical oxidation as a result of the airing of the sediments (Smettan, 1998: 61). The pollen survival rate is considerably better in waterlogged conditions in pond and lake sediments, but these are lacking in Malta. Preservation may also be poor in marine deposits (Lowe & Walker, 1984) although several authors (e.g. Jahns et al., 1998: 277-288; Sun & Li, 1999: 227) have achieved good results with pollen analyses on marine deposits at various locations. Nonetheless, the settling of some pollen types through the water column in the sediments of Marsa Core 1 may necessitate a very calm environment.19

Carroll identified the pollen grains by reference to Moore et al. (1991), Jansonius & McGregor (1996) and University in-house pictorial resources at the University of Huddersfield. Additionally, a photographic reference library was created from material identified by Chris O. Hunt and from type slides prepared by John Corr at the University of Leeds from material supplied by the University of Malta.

7.3.2.1. Material and methods Often, a non-negligible fraction of a pollen assemblage may consist of redeposited pollen grains. While in many regions these may often be easily identified due to their incompatible stratigraphic ranges (e.g. Tertiary pollen in a Late Pleistocene sediment), in other cases the reworked pollen may not appear obviously older than the enclosing deposit, and its presence may be more difficult to detect (Campbell, 1999: 246).

The pollen analyses of various samples from Marsa Core 1 were done by Frank Carroll, who kindly made the results available for the present study. Carroll’s palynology preparation followed Hunt (1985). This method is considered safe, efficient and inexpensive. Furthermore, the sophisticated safety equipment was not available for the more widely used hydrofluoric (HF) acid digestion and funds to pay for HF preparation at an outside source were insufficient. Applying Hunt’s technique, approximately 15g of sample material was placed in a 200cm³ beaker, and boiled on a hotplate for 20-30 minutes with a 10% solution of potassium hydroxide (KOH) and sodium pyrophosphate. After cooling, the sample was decanted through a 106 micron metal sieve onto a clock-glass, where it was ‘swirled’ to allow the sand and silt to settle; the remaining suspension was passed through a 7 micron nylon mesh. After filtering through the mesh, 3M hydrochloric acid was continually added until calcareous material was

A re-investigation of some of the pollen slides under a fluorescent microscope at Queens University, Belfast, revealed the presence of recycled and intrusive pollen among the assemblages. This led to a complete reinvestigation of all pollen slides prepared by Carroll, who determined intrusive pollen by a stained pale pink colour and containing protoplasm, while recycled material stained more purplish, often very pale. The recycled material is an important indication of re-deposition of sediments, while the source of the intrusive pollen may perhaps have been the water from the water-bowser that was used in the process of retrieving the cores. Both the recycled and intrusive pollen are excluded from the pollen diagrams.

19

For example, the pollen of Pinus (pine) are equipped with airsacs, which facilitate their dispersal by wind. These airsacs also help the pollen to float on water and may prevent it from sinking.

95

A considerable number of different tree taxa were detected by Carroll in the pollen slides, but most of the singular occurrences are likely to be due to inblowing from nearby Sicily and/or North Africa. The pine Pinus is by far the most common tree species, followed by the cypress family Cupressaceae. Also prominently present is the olive family Oleaceae and the myrtle family Myrtaceae, while the presence of oak Quercus is overall very low. Pinus, which dominates the overall pollen spectrum in the lowermost 2m, experiences a drastic decline from 950cm upwards, from which it never recovers (Figure 7.3.1.). This dominance, however, may perhaps only be relative due to the comparatively lower susceptibility of pine pollen to oxidation (Havinga, 1964: 621).

7.3.2.2. Results The results are presented in diagrams of pollen percentages against depth and were plotted with Tilia and Tiliagraph (© Grimm, 2004) by Frank Carroll, and the counts are expressed as a percentage of the total pollen sum. The overall presence of pollen grains was at times very low and the pollen sum per analysed slide was often much lower than 100 grains, particularly towards the base of the core, which makes statistics unreliable for these levels. Therefore, in order to achieve a more meaningful graph, the samples were successively bulked-up whenever necessary, until at least 100 pollen grains were reached. However, the very low pollen counts are likely to be due to differential redeposition and preservation, particularly in the upper 100cm of the core, as well as dilution in the lower part of the core. As pollen grains are also affected by various processes between dehiscence and recovery, this may result in important biases in the pollen assemblages (Campbell, 1999: 245).

Compared to the trees, shrubs form only a minor group, and here the Rosaceae family dominates. Herbaceous plants form by far the largest plant group, and they dominate the picture from around 4840 cal. BC onwards, although several herbs occur throughout the column (Figure 7.3.2.). Most prominent is the subfamily Lactucae (lettuce group) of the family Asteraceae, particularly in the upper 2m of the core. Also well represented are the wormwood Artemisia, the composite family Asteraceae, the goose-foot family Chenopodiaceae, the carnation family Caryophyllaceae,

Preservation was much better in the waterlogged silty grey sediments, particularly between 300cm and 600cm of Marsa Core 1. The amount of damaged pollen varies broadly parallel with the total amount of pollen, while the recycled pollen appears to be mainly present in the oxygenated sediments in the lowermost 4m of the core as well as in the upper 3m.

530-660 cal. AD

800-500 cal. BC

4690-4460 cal. BC

21750-20800 cal. BC 4840-4610 cal. BC

> 44400 BP 22350-21750 cal.BC

Figure 7.3.1.: Marsa Core 1, Pollen Analysis (except herbaceous taxa), count expressed as a percentage of the total pollen sum. Pollen sums have been bulked up to reach at least 100. The depth axis on the left refers to the depths as used in the present work, while the depth axis on the right refers to the depths as used by Frank Carroll and is given to enable comparison of the present work with that of Carroll. Values for 'pollen sum', 'damaged pollen' and 'recycled pollen' are absolute numbers, not percentages. Data source and Tiliagraph: Frank Carroll.

96

530-660 cal. AD

800-500 cal. BC

4690-4460 cal. BC

21750-20800 cal. BC 4840-4610 cal. BC

> 44400 BP 22350-21750 cal.BC

Figure 7.3.2.: Marsa Core 1, Pollen analysis of the predominantly herbaceous taxa, count expressed as a percentage of the total pollen sum. Pollen sums have been bulked up to reach at least 100. The depth axis on the left refers to the depths as used in the present work, while the depth axis on the right refers to the depths as used by Frank Carroll and is given to enable comparison of the present work with that of Carroll. Data source and Tiliagraph: Frank Carroll.

the sedge family Cyperaceae and the true grass family Poaceae. There is also pollen of several plants that are associated with cultivation (e.g. Agrostemma, Papaver and Silene), but these occur in low numbers (see Figure 7.3.2.).

7.3.3. Plant Macro-Remains The majority of the samples analysed from Marsa Core 1 also yielded plant macrofossil remains in varying amounts. These consisted of charcoal, dried plant fragments, decomposed remains and seeds and were sorted out during the general sorting process of the molluscs. In many cases the preservation of the plant macrofossils was surprisingly good and often allowed determination to species level. The macrofossils consisted mainly of seeds, spikelets, leaves, stems and awns. In other cases, the fossils were burnt, heavily fragmented or in an advanced stage of decomposition. As with charcoal (see below), their presence in the core is likely to be indicative of the immediate environment, but they can also represent downwash from the watershed into the basin during heavy rainfall events.

Cultivated plants (see Figure 7.3.1.) consist mainly of cereals, and these occur as from 975cm upwards, but their percentage is very low. Agricultural activity appears to increase as from 940cm upwards, but experiences variations and the pollen percentage never exceeds 12%. There may be a correlation between the occurrence of cereal and the decrease in arboreal pollen (see Figure 7.3.1.). The flax Linum is not included here, as it may be a native to the Maltese Islands, whose use may have been exploited, however, at a considerably later date, as suggested by the results (see Figure 7.3.2.). Wetland species and ferns are also present, more or less throughout the core, and particularly the occurrence of wetland species coincides largely with the presence of freshwater shells (compare Figures 7.1.2. and 7.3.2.).

Identification was done by the present author with the help of reference books (Brouwer & Stählin (1955), Haslam et al. (1977), Hanf (1982), Schoch et al. (1988), Davis (1993) and Martin & Barkley (2000)) and through comparison with recent specimens. The Phoenician juniper, Juniperus phoenicea remains were kindly identified by José Carrión of Murcia University, while several seeds were kindly identified by Dr Charles Turner of Cambridge University. A number of macrofossil plant

The plants represented in the pollen spectra stem from a variety of habitats and soil preferences. This suggests also the regional aspect of the pollen sources that were united within the sampled sediments of Marsa Core 1.

97

remains were unidentifiable due to bad states of preservation.

Poaceae, would indicate that significant open areas of grassland were also present (Huntley et al., 1999: 951). However, the pollen concentration is very low in the lowermost 2m of the core and it thus cannot be excluded that the dominance of pine pollen may be an effect of differential decomposition (Havinga, 1984; RossignolStrick & Planchais, 1989). Selective destruction of pollen has been found to occur already at relatively low levels of oxidation, and Pinus, a notably resistant pollen that is shed in millions from the tree, may here, in fact, be responsible for taphonomic skewing (Hopkins & McCarthy, 2002: 167).

The overall number of macrofossil plant remains is low, and as such, not comparable to the other data sets. Nonetheless, the presence of these plants is very interesting. Appendix III lists the remains as found and not standardised. 7.3.3.1. Results The determined plant species complement the pollen spectra very well and add a number of species and genera that are not reflected in the pollen record. Apart from the leaves of the Phoenician juniper,20 the other woody taxa detected among the plant macro-remains were Ficus carica (fig), represented by a comparatively high number of seeds; Olea europaea (olive), of which fragmented kernels were found and a member of the prune family. The occurrence of trees is comparatively more common in the lowermost 4m of the core, particularly from the base until 975cm, which seems to agree with the pollen record (compare Figure 7.3.1.).

There may have been woodland, which would have consisted mainly of pine trees (Pinus) and a few oak trees (Quercus sp.). Trees from the cypress family (Cupressaceae) and fig (Ficus carica), more indicative for scrub or open woodland (Haslam et al., 1977: 14; Huntley et al., 1999: 946), occur concurrently as indicated by the pollen diagram combined with the macrobotanic remains (Figure 7.3.1. and Appendix III). The presence of these trees seems to decline very rapidly from 960cm upwards, while palms (Palmae) and the olive (Olea), suggestive of open spaces under a summerdrought rainfall regime (Carrión & van Geel, 1999: 213) increase, albeit in very low numbers.

The largest plant group are the herbaceous plants and of these, the majority are associated with waste places and cultivated land. Of non-arboreal plants, only the vine Vitis vinifera and sulla Hedysarum coronarium left their traces as cultivars in the samples. Sulla occurs only once, but was found in a deposit datable to the Neolithic. Fragments of grape seeds were found in a sample that considerably pre-dates the Roman period (Sample 96, Late Bronze Age/Phoenician), but also occurred in Roman Period levels, as well as in the Late Middle Ages (see Appendix III).

There may perhaps be a link between the first occurrence, and hence, cultivation of cereals (Cerealia) and the decrease in trees between 1075cm and 940cm, although again, the absolute number of pollen grains found was very few. Percentage-wise, the amount of cereals present in the pollen slides remains below 10% of the total pollen sum (except for the 11% peak between 455cm and 430cm) and there broadly appears to be no clear correlation between any increases or decreases in cereal pollen and arboreal pollen beyond 940cm (see Figure 7.3.1.).

Hygrophytes and hydrophytes form the second largest plant group, and their presence further confirms the presence of a perennial freshwater stream, which formed a brackish water body at Marsa.

Of the predominantly woody taxa, the pollen of Abies, possibly Acacia, Alnus, Betula, Carpinus betulus, Fagus, Picea, Tilia and Daphne are likely to have been inblown from Sicily and/or North Africa or carried with currents to the Maltese shores. The occurrence of these exotic pollens is very infrequent and when they occur, they do so in very low numbers, which may be indicative of the distance across which they have been transported. Furthermore, several of these species survive only in pockets in Sicily, which have no ecological parallel in Malta, like e.g. Abies and Betula (Sadori & Narcisi, 2001: 657).

7.3.4. Analysis of vegetation change throughout the core In spite of the paucity of pollen remains from generally oxygenated sediments at the lowermost 4m of the core and towards the top of the core, bulking-up of successive pollen slides reveals a very interesting picture, which may, however, be affected by taphonomic biases. The most prominent feature of the pollen analysis, and supplemented by the macrobotanical remains, appears to be the change from a predominantly wooded environment to a predominantly open country and steppe vegetation, as indicated in the changing percentages between arboreal (including trees and shrubs) and non-arboreal pollen in the lower part of Marsa Core 1 up to 730cm (see Figure 7.3.1.). Only once, at 960cm, is the value for arboreal pollen greater than 90%. Otherwise, the persistent values above 15% for herbaceous pollen taxa, dominated by

The presence of olive (Olea) may seem to fluctuate throughout the core, yet interestingly, the highest amount of olive pollen has not been found in samples corresponding to the Roman Period, but in samples datable to the Neolithic and to the Phoenician/Punic period. Similarly, grape seed fragments of vine (Vitis vinifera) were found at levels corresponding to the Late Bronze Age and the Phoenician/Punic periods, indicating that viticulture may have started in Malta significantly prior to the Roman period (see Appendix III). Also,

20 This very drought resistant tree no longer occurs in the Maltese Islands (see Haslam et al., 1977).

98

pollen of the walnut Juglans was detected at levels corresponding to the Late Bronze Age and Phoenician/Punic Period, albeit in very small numbers. The occurrence of Juglans in a possible Late Bronze Age context makes its occurrence in Malta perhaps concurrent with its appearance in neighbouring Sicily, where its cultivation could be dated back to about 950-400 cal. BC BP (Sadori & Narcisi, 2001: 669).

As a number of problems exist with the current nomenclature, the terminology used herein will follow Goldberg, 1985, and Jones et al., 1997. ‘Charcoal’ refers to “any black-coloured plant-derived material that has had its chemical composition and ultrastructure significantly changed as a result of heating in a fire. It retains recognisable anatomic structure of the parent plant. When carried with smoke emissions, the particles produced in fuel by fire are referred to as “small charcoal particles”. The particle size here ranges from ca 1 micron to several mm”

Ferns are present throughout the core in varying amounts, as are wetland species. Particularly the wetland species increase between 300cm and 200cm (see Figure 7.3.1.). A large variety of herbaceous plants occur throughout the length of Marsa Core 1. Of these, the presence of the wormwood Artemisia is generally considered as an open country indicator (e.g. Schwab et al. 2004: 1730), and this is present in varying amounts throughout the core, and in small amounts already from the base of the core upwards. Similarly, the presence of Rumex, also indicating open woodland (Sadori et al..2004: 12) is noted at the base of the core (see Figure 7.3.2.). Several herbs associated with agriculture like Papaver, Agrostemma, and Silene, occur irregularly, but are generally confined to the middle section of the core. In the upper 2m of the core, the percentage values of herbaceous plants generally exceeds 90%.

(Jones et al., 1997: 20). The term ‘black carbon’ is generally employed to describe any product of incomplete combustion of fossil fuels and vegetation; black carbon exists as a continuum from partly charred plant material21 through char22 and charcoal to graphite and soot particles,23 with no general agreement on clearcut boundaries (Seiler & Crutzen, 1980). Thus, the formation of black carbon can occur in two fundamentally different ways: from recondensation of volatiles to highly graphitised soot, and from solid residues that form charcoal. Any form of black carbon is purely terrestrial in origin, and is relatively inert (Schmidt & Noack, 2000: 777), although it may be subject to fragmentation, especially when transported by water (Patterson et al., 1987: 3). Dating of soil charcoal demonstrates that charcoal is preserved in soils and sediments for tens of millennia (Dickens et al., 2004a: 338), while black carbon residues are part of many coals that date back to the Devonian (Taylor et al. 1998).

The peaks in occurrence of several herbaceous plants between 860cm – 840cm and between 690cm - 310cm may be related to a better preservation of pollen grains due to anoxic conditions in the waterlogged deposits there. This is also indicated by the significantly higher pollen concentrations, which on two occasions even exceed 900 pollen grains per slide (see Figure 7.3.1.). The amount of damaged pollen seems to agree with the changing depositional environments. Studies have shown that pollen generally experiences comparatively little damage when transported with dry sediments, but gets more easily damaged in waterlogged conditions (e.g. Havinga, 1984: 541-558; Dupont, 1998: 26).

7.4.2. General considerations The production and the amount of black carbon produced depend on the fire temperature (Umbanhowar & McGrath, 1998: 341), weather conditions and vegetation characteristics, as these determine the combustion efficiency and the rate of fuel consumption. Studies have shown that a savannah landscape may blaze fiercely annually, but consumes only 10cm) overlies this deposit and it may have contributed to

8.3. Section 2: From the Neolithic to Byzantine Malta The sediments between 730cm and 320cm comprise nearly 5000 years of Maltese history, as indicated by the radiocarbon dates. The predominantly grey sediments in this section also represent the brackish lagoon phase of what is today the Marsa plain and it also appears to be characterised by several large-scale events, despite the

Approx. water depth at around 1355 cal.BP

Possible water depth at around 2000 cal. BP Approx. water depth at around 2600 cal. BP

Waterlogging possibly around 6525 cal. BP

Figure 8.2.: Holocene rise of sea levels according to data from Sicily and nearby Italy, and transposed on Marsa Core 1. Note that the graph may represent an idealised rising of the levels, particularly in the lower levels. Source of sea level data: Lambeck et al., 2004. 1 For example, the occurrence of a single boulder on a muddy bottom results in a different sedimentation behaviour than if the same muddy bottom was homogenously covered by boulders.

107

the anoxic conditions that rendered the sediment nearly black. The large cobble/boulder may be the result of an extreme event, but the nature of this event cannot be determined from the evidence of the core. The calculated date range up to 710cm would place this part in the Late Neolithic/Early Temple Period. Macroscopic charcoal was nearly absent, and pollen and plant macro-remains were very few. While the cultivation of cereals appears to have increased, so has the percentage of arboreal pollen. Ferns, on the other hand, decrease, which may perhaps be indicative of an aridification of the climate. This is further suggested by a substantial decrease in the absolute pollen sum (see Figures 8.3.). The land snail assemblage in this part of the core would indicate initially an open country/garrigue environment that may have favoured an increased species richness and abundance as reflected between 730cm and 720cm. This appears to be followed by a drastic decrease in the diversity, and the increased abundance of the opportunistic and ubiquitous species Cochlicella acuta may point to harsher environmental conditions.

silt- and clay sized particles increases gradually, which may possibly indicate calmer settling conditions, perhaps also as a result of still rising sea levels. Calmer conditions are also implied by the comparatively high pollen sum and rising number of brackish water and marine molluscs, while the very low number of land snails may perhaps indicate a diminished influx of terrigenous sediments. The land snail species found here would point to less hostile conditions than before, as the overall diversity increases and the presence of the opportunistic Cochlicella acuta decreases. The presence of Granopupa granum, Muticaria macrostoma and Trochoidea spratti suggests an open country/karstland environment with garrigue vegetation (Giusti et al., 1995: 221; 354; 404). Charcoal particles were remarkably few. Interestingly, pollen of Oleaceae increased, possibly an indication of olive cultivation, while cereal cultivation appears to have decreased. After an initial drop in arboreal pollen, their number again increased slightly by 680cm, perhaps as a response to a lessening of arid conditions, as also suggested by a rise in Pteropsida pollen (ferns).

Thus, a possibly reduced plant cover may have led to increased rates of run-off and erosion, which could in turn have led to the deposition of large cobbles/boulders in the fan catchment. Similar effects have been observed elsewhere (e.g. Went, 2005: 693)

This relatively favourable picture is drastically altered by one or a series of extreme sedimentation events that led to the accumulation of 75cm of material. This 75cm section of the core consisted of boulders > 25cm at the bottom at 675cm, overlain by < 50cm of very coarse and wet material that could not be retained by the mechanical corer due to compression difficulties (see above, Chapter IV). In the absence of any examinable material from this event, it is difficult to speculate about its cause. A very big rainfall event and a turbidite may have led to this

The grey sediments overlying the large cobble/boulder from 700cm to 675cm are likely to correspond to the end of the Neolithic Period and to the Temple Period until the Tarxien Phase (see Figure 8.1.). The sedimentation of

Pinus 0 10

20

30

40

OLEACEAE 50

60

0

10

20

30

CUPRESSACEAE 40 0 2

4

6 8 10 12 14 16 18

Linum 0

1

2

Pte ropsida 3

4 0

10

20

CEREALIA 30 0

10

20

30

40

0

1

2

3

depth (m)

4

5

6

7

8

9

10

PANGAEA/PanPlot

max.: 10.95 m

absolute values MC1 pollen2.txt - 24.06.2006 15:54 h

Figure 8.3.: Marsa Core 1, pollen analysis of selected taxa, counts expressed in abslute number. Note that the X-axis does not have the same scale in all cases. Source of data: Frank Carroll.

108

deposition. According to the calculated dates, this might perhaps have coincided with the rise of megalithic constructions and it is tempting to speculate about a possible relationship between these two occurrences. Perhaps the construction of the first megalithic buildings may have been indicative of socio-economic activities that perhaps may have led directly or indirectly to a destabilised soil or ground cover, which, during a major rainfall event may then have resulted in this excessive sedimentation. But it may also have been the other way round: the formerly widespread stone/mud-brick huts and shrines as found at Skorba may perhaps have continuously crumbled away during extreme wet weather events, resulting in considerable losses that may have often endangered survival. The constructions of larger buildings out of much larger stones that withstand extreme wind and rainstorms much better than simple huts, whether for shelter and/or for veneration may have been a logical response.

predominantly brackish water molluscs indicates a comparatively undisturbed waterlogged environment. The resolution of the pollen record in this part is low (four samples, irregularly spaced), but still indicates a fluctuating spectrum, although the fluctuations are generally minor. Cereal production seems to go up and down, Oleaceae (olive) seems to recover from the previous low and climbs steadily, while Pinus (pine) pollen decreases to zero values (Figure 7.3.1.). Ferns also decrease after an initial rise, perhaps indicating more arid conditions. The land snail assemblage also reflects a predominantly open country and karstland environment. A noteworthy increase in freshwater molluscs may then perhaps be indicative of relatively more severe rainfall events. The presence of three now extinct freshwater molluscs (Planorbis planorbis, Bulinus cf. truncatus and Carychium cf. schlickumi) suggests a perennial freshwater body fed by a spring (Giusti et al., 1995: 162; 180; 198). An increased presence of burrowing land snail species (Ferussaccidae) may point to more pronounced erosion events, perhaps as a result of increased precipitation (see Figure 7.1.5.). Macroscopic charcoal is present in higher amounts than in the previous 250cm of the core, which may suggest increased firing in the vicinity (Figure 7.4.1.). Plant macro-remains from this part of the core reveal the presence of the vine Vitis at 540cm. Pollen of vine has been found in Sicily already in Early Holocene deposits, but the beginning of its cultivation there is hard to assess (Sadori & Narcisi, 2001: 669). Vine may have been introduced to Malta from Sicily during the Late Bronze Age, possibly in the Borg in-Nadur phase.

A radiocarbon date at 600cm revealed that the submitted organic material consisted of contaminated recent matter (see above, Chapter V). Which period or periods the sediments between 600cm and 515cm exactly cover is thus largely a matter of speculation. A radiocarbon date at 510cm revealed a range between 800 – 500 BC (94.2% probability), which would make the sediments below possibly of Bronze Age Period age and perhaps also the end of the Temple Period, if the sedimentation continued unabated after the previous extreme event. Molluscs, predominantly brackish water species, appear to slowly re-colonise the lagoon from 600cm onwards. Their comparatively low number may perhaps also be an indication of a relatively rapid deposition, as also suggested by the poorly sorted sediments up to 585cm. Land snails occur in low numbers only, and the species again indicate predominantly a karstland and steppic environment (presence of Granopupa granum, Muticaria macrostoma and Trochoidea spratti). The pollen spectra shows slight variations, notably a drastic decrease in Oleaceae (olive) and Cupressaceae (cypress) pollen, and other heliophytic open country indicators like Artemisia, Plantago and Poaceae (grasses). An increase in pine Pinus and oak Quercus pollen points to an increase in woody taxa, while the percentage of cereal pollen decreased (Figures 7.3.2.). This part of the core may perhaps belong to the end of the Temple Period (see Figure 8.1.), and thus could point to a temporary abandonment or at least a possible decrease in population, however, without a radiocarbon date nothing can be said. The amount of macroscopic charcoal is too unspectacular to be noteworthy.

In Sicily, at Lago di Pergusa, the first appearance of the cultivated walnut Juglans is dated to about 950-400 cal. BC (ibid.), and its introduction to Malta, possibly in the early Phoenician/Punic phase is likely to have happened roughly at the same time, as indicated by the presence of Juglans pollen at 535cm. This may perhaps indicate the beginning or early phase of the local Phoenician/Punic Period, as this tree is a native of the Near East (Athar & Mahmood, 2005: 41) and was diffused in the Mediterranean by the Phoenicians. During this period, the incidence of fire appears to increase, with a first notable peak of macroscopic charcoal at 525cm, perhaps as a result of increased anthropogenic activity. The influence of the perennial freshwater body appears to diminish considerably at the coring site of Marsa Core 1. The now extinct species Bulinus cf. truncatus and Carychium cf. schlickumi vanish from the molluscan assemblages. However, the perennial stream continues to flow as evidenced by the occurrence of other freshwater molluscs, most notably by Pseudamnicola moussonii; it is possible that the stream may have experienced such a reduction of its former size that it was no longer suitable for these now extinct species. A similar observation has been made at the Quaternary deposit of Bahrija Valley (Giusti et al., 1995: 200). The reduction of the stream may be the result of anthropogenic action, perhaps diversion for irrigation purposes, rather than the result of a prolonged drought that may have dried up the stream.

The sediments between 535cm and perhaps beyond 585cm may possibly belong to the Bronze Age Period in Malta. The particle size distribution (Figure 6.1.) in this part of the core indicates little stability in the depositional environment of the shallow Marsa lagoon, with pebblesized stones finding their way to the location of the borehole of Marsa Core 1. Nonetheless, in times of quiet sedimentation, the accumulation of large numbers of 109

The land snail assemblage, or occasional lack thereof, in this part of the core may perhaps support this suggestion: an increased vegetation cover, indicated by the leaf litter species Truncatellina callicratis, may also have prevented the downwashing of many landsnails onto the coring site of Marsa Core 1, which would perhaps explain the paucity of their remains (Figure 7.1.3. and Appendix). A radiocarbon date at 515cm revealed a range between 800–420 BC (95.4% probability), with an emphasis on the range between 800-500 BC (94.2% probability. Thus, sediments corresponding to the Phoenician/Punic Period may extend up to perhaps around 470cm. In this part of the section, the percentage of cereal pollen appears to first decrease, but then rises steadily. When considering the absolute pollen count (Figure 8.3.), however, the number of cereal pollen reaches unprecedented high levels. A similarly diverging picture emerges too from comparing the olive pollen in both graphs, which, in Figure 8.3. also reaches its absolute peak at 510cm.2 The high amounts of these cultivars may well reflect an increased agricultural activity, as suggested by the non-marine molluscan assemblages above, and as also indicated by the distribution of Phoenician, Punic tombs and presence of several farmsteads in the Maltese countryside that were then later used during the Roman period (see above, Chapter III). This development may perhaps be a reflection of the favourable climatic conditions that then prevailed in the Central Mediterranean (Faust et al., 2004: 1771 and see above, Chapter II). The Cupressaceae seem to have their highest abundance during this period (see Figure 8.3.) and may well include the sandarac gum tree Tetraclinis articulata. However, it is impossible to differentiate Cupressaceae pollen down to species level (Carrión, personal communication: 2006). Despite a possible increase in agricultural activity and an increased vegetation cover, macroscopic charcoal remains do not point to increased fire activity.

reason for the fire that disrupts this potentially stable environment may perhaps not be climatic, as the low sedimentation rates do not increase immediately after the event at 475cm. However, the immediately resulting reduced plant cover may have led to the notable increase in land snails that were washed down onto the coring site of Marsa Core 1. There is also a remarkable increase in freshwater molluscs at 475cm, perhaps as a result of an increased freshwater flow. From the calculated radiocarbon range, this event or maybe series of events may have occurred between ca. 500 BC and ca. 180 BC, with an estimated 85% of this event occurring between ca. 300 –180 BC (see Figure 8.1.). Perhaps this fire may be connected with the historic raid of the islands during the First Punic War by the Roman general Attilius Regulus at around 255 BC, who ravaged, burnt and laid waste the whole island. This event is mentioned by Naevius (IV, fr. 32) and Orosius (4.8.5.), but archaeological evidence for this presumed widespread devastation by fire has so far not been detected in Malta (Bonanno, 2005: 79). The vast amounts of macroscopic charcoal would point to a local fire in the immediate vicinity of the lagoon in Marsa, but how widespread the fire may have been cannot be assessed.3 However, if Marsa was a key harbour in those days, then an attack centred on Marsa would be plausible. If the former reduction of the perennial stream was anthropogenic and perhaps for irrigation purposes, its possible destruction, as suggested by the sudden increase in freshwater molluscs found in the corresponding sample, may have been short-lived. Agricultural activity may have been resumed soon afterwards as suggested by the subsequent drastic decrease in land snails and freshwater molluscs (see Figure 7.1.3. and Appendix II). A period of relative stability appears to follow this major fire event up to 425cm. At 460cm there may be evidence of flax cultivation in the vicinity of Marsa, as indicated by Linum in the pollen diagram. Flax already occurred very sporadically in Neolithic levels at Marsa (see Figure 7.3.2.), but as Linaceae occur naturally in the Maltese Islands, its anthropogenic use is difficult to assess (Haslam et al., 1971: 172-4). The overall absolute amounts of Linum pollen are very low, which is somewhat surprising, considering the importance Cicero attaches to Melitensis, the Maltese cloth presumed to have been woven out of flax in Malta in the first century BC. According to the calculated sedimentation rates and extrapolated dates, 460cm would correspond mainly to the 3rd and 2nd centuries BC, but as Linum has not been found in the analysed pollen slides between 435cm and 330cm, flax retting may have been a comparatively shortlived industry in Roman Malta, where the value of the woven cloth may perhaps also have been derived from the scarcity of the flax itself.

The period of enhanced agricultural activity and productivity may perhaps have led to greater economic activity. Following Lambeck et al.’s (2004) sea level curve (see Figure 8.2.), economic activity in the Grand Harbour may now have been favoured by the still continuously rising sea levels, which may have turned the formerly very shallow lagoon into a more navigable water body with depths exceeding 1.3 m. A drastic change occurs at 475cm, where the amount of charcoal reaches a record peak with more than 700mm3 (see Figure 7.4.1.). The pollen spectrum is here of a low resolution, the nearest analyses have been made at 485cm and 460cm. The erosion rates may be comparatively low around this phase, as indicated by the deposition of predominantly muddy sediments (see Figure 6.1.). The

Cereal cultivation appears to continue at unabatedly high levels until at least around the 2nd century BC – 1st century AD (at 435cm), as indicated by the extrapolated dates. Plant macro-remains of olive Olea europaea and

2 These discrepancies arise from the percentage calculations, in which each type directly influences all others within the sum. However, as Carroll employed total counting techniques of analysis, it is also possible to express the pollen data in terms of numbers/sample. Thus, the value for any taxon is independent of all others and one can follow trends in the abundance of the taxa in time (see Moore et al., 1991: 170).

3

110

Microscopic charcoal has not been analysed at this level.

vine Vitis were found for the last time at 435cm, and may perhaps substantiate the possible last peak of agricultural activity and cultivation in the Roman Period.

and 340cm may point to an increased vegetation cover that prevented the shells from being washed onto the coring site of Marsa Core 1. Although the analysed pollen is again considerably spaced out (nearest records are at 410cm and 355cm), the pollen data at 355cm indicate an increase in predominantly woody taxa and a marked decrease in cereal pollen, which may perhaps be a reflection of a decreased population and/or a marked decrease in agricultural and economic activity (see Figure 3.9.). The disappearance of a usable harbour at Marsa may have also contributed to this economic decline.

Then, perhaps in connection with another major fire event that left its trace in the sediments at 420cm (around 1st to 2nd century AD) there seems to be a drastic decrease in the absolute amount of cereal pollen from around the 2nd century AD. A progressive decrease in the vegetation cover may, however, have preceded this event as is suggested by a drastic increase in land snails and freshwater molluscs at 425cm (see Figure 7.1.3. and Appendix II). The fire would have aggravated the lack of vegetation and after this event, the recovery from this loss of vegetation cover appears to take longer than the presumably equivalent event in the 3rd century BC (see above), perhaps as a result of less favourable climatic conditions, but perhaps there may also have been a decrease in the overall population. Cereal pollen in Marsa Core 1 then reached an all-time low around 3rd – 5th centuries AD, before attaining high levels again around the 6th century AD. Interestingly, this same trend has also been noted by Bruno (2004) in her economic analysis based on amphorae analyses (compare Figure 3.9.).

A radiocarbon date at 320cm gave a Byzantine date ranging between 530 AD and 660 AD (95.4% probability). Hence, the immediately preceding sediments (between 335cm and 325cm) may belong to the 5th and 6th centuries AD, when the Vandals spread across the Mediterranean. Although here the archaeological evidence is very scanty (see above, Chapter III), the pollen analysis appears to indicate an increase in olive and cereal pollen, perhaps reflecting an increase in population and/or economic activity (see Figure 7.3.1.). From the total pollen amount, the increase appears substantial, particularly for cereals (see Figure 8.3.), while the percentage distribution of both olive and cereal show rather minor increases at 330cm. Concurrent with the relatively high amount of cereal pollen are unprecedented high amounts of the corn cockle Agrostemma, and goosefoot pollen (Chenopodiaceae), both associated with cultivation (Haslam et al., 1971: 27ff, 52).

Between 390cm and 360cm, which corresponds to some time during the late 3rd or 4th century AD, another event of rapid sedimentation seems to characterise the stratigraphy of Marsa Core 1; again, the sediments were too coarse and wet to be retained by the mechanical corer. Thus, in the absence of a physical examination of the deposited material, it is well nigh impossible to determine if these 30cm of material were of terrigenous or marine origin. During the 4th century AD, several very strong earthquakes occurred,4 and it is very possible that they may have had a similar strong impact in Malta as they did in nearby Sicily and Libya. Thus, the lost material may well have consisted of a tsunami washover deposit. Alternatively, a decrease in agricultural activity, perhaps in connection with an extreme weather event may have led to the erosion event, resulting in an increased input of sediment. However, the absence of any land snail species in the deposits immediately prior and after this purported event does not support this interpretation (see Figure 7.1.2.). Whatever the origin of this material, its deposition may have made use of the lagoon at Marsa as a harbour increasingly difficult because of the shallowing waters (see also Figure 8.2.). This seems also to be indicated by the archaeological remains in the form of warehouses that were found along Racecourse Street in Marsa (see above, Chapter III), which were in use since the 4th century BC, but which were abandoned in the 4th century AD (Bonanno, 2005: 239). Hence, it is possible that this sedimentation event may be directly related to the lack of archaeological material from Marsa in later times. The absence of any land snail species found between 395cm

Macroscopic charcoal remains were unspectacular. The markedly biggest change, however, seems to occur at around 330cm, when the molluscan assemblages change considerably and indicate an unprecedented increase in freshwater molluscs, combined with a clear decrease in brackish water and marine molluscs. Land snails also increase, but as the fragments found are predominantly of ubiquitous species, little may be said about the environment. The overall number of shells, however, decreases dramatically, possibly a sign of rapid deposition of sediments through erosion that caused a mudflow. Erosion is perhaps also indicated by the presence of the burrowing land snail Ceciliodes acicula, which may have found its way into the samples through the removal of deeper sediments (see Figure 7.1.5.). The cause for this erosion may perhaps be a combination of increased agricultural activity with increased precipitation, as suggested by the increase in ferns (Pteropsida). 8.4. Section 3: From Byzantine Malta to the 15th century AD The remaining sediments of Marsa Core 1, between 320cm and 50cm,5 contain less than 1000 years of history

4

For example, a series of earthquakes between 361-363 AD that destroyed many cities in Libya and Sicily, and the earthquake of 21 July 365 AD, which occurred in Crete and its resulting tsunami created enormous damage also in Sicily (Guidoboni et al., 1994: 260, 267-274 and see above, Chapter II).

5 The 50cm sediments retrieved from the first tube had been compressed to 30cm length by the mechanical corer. Furthermore, the uppermost 15cm were from the disturbed ground. Hence, the uppermost 50cm are not considered here.

111

and have accumulated as a result of continuous erosion, possibly coupled with a much slower rise in sea levels than in the early Holocene. This has progressively turned the former shallow lagoon environment into marshland.

drastic decrease in the land snail assemblages found between 285cm and 265cm, as again an increased vegetation cover may have prevented the shells from being washed onto the coring site of Marsa Core 1. Only present in a fragmented state, the land snail remains cannot be more closer determined than ‘Heliciid’, and as such contribute only little to the environmental reconstruction. Pollen analysis from 280cm reveals that cereal pollen is nearly absent from the spectrum, while flax is still present but in very low amounts. Although the 15th century writer Al-Himyari mentions the former presence of pines, junipers and olive tree, the pollen spectra show a notable absence of these trees, perhaps already as a result of the visits by shipbuilders and other wood traders, who could have felled the trees indiscriminately, perhaps for resale at Scicli in Sicily, as indicated by El-Idrisi (see above, Chapter III). Thus, despite a possible phase of depopulation, woody taxa appear not to have increased on the islands.

A radiocarbon date at 320cm with a range between 530 AD and 660 AD (95.4% probability) fixes the base of this section, while the clinker layer, which was laid down in 1869 (see above) is the upper limit. However, this upper date is not without problems: it is very possible that an unknown amount of sediments was removed – or added – before the clinker layer was put down. This is also suggested by the irregular occurrence of rubble and limestone chippings in the section (see Figures 1.4.1.– 3.). Furthermore, sediments were lost from the mechanical corer due to compression problems at the level of the water table (15cm lost between 130cm and 145cm), or compressed together (between 30cm and 50cm). Extrapolation of dates for these sediments should thus be done with caution and will here be used mainly as a broad guidance.

Climatically, there was a cool and dry period that led to increased erosion in the Central Mediterranean at around 1000 BP (see above, Chapter II), and this may have also been the cause of increased sedimentation rates and adverse conditions for tree regeneration. This was followed by the Medieval Warm Period, which, in North Tunisia resulted in rising temperatures, stable sedimentation rates and more humidity around 900-700 BP (Faust et al., 2004: 1776). In Malta, this time range would correspond to the Post Arab Period (1091 – 1282), and in Marsa Core 1 perhaps to the sediments around 245cm – 225cm, as indicated by stable sedimentation of silts and clays (see Figure 6.1.) and comparatively few land snails in the molluscan assemblages. The 1169 earthquake, which caused massive destruction in Sicily (see above, Chapter II) does not appear to have left a detectable signature in the sediments of Marsa Core 1, but the absence of a reliable absolute date makes it hazardous to speculate about it. At 230cm, the nearest point in the core where pollen has been analysed, woody taxa increase again, as do cereals and flax, as indicated by the absolute pollen graph (see Figure 8.3.), perhaps as a response to the more favourable climatic conditions. An increase in the vegetation cover is furthermore suggested by a comparatively low amount of land snails in the molluscan assemblage at 230cm. From 220cm upwards the land snail assemblages become more varied and abundant. The species found indicate again the presence of an open country with karstland features and steppic vegetation, but humid conditions and perhaps the vicinity of a wetland margin are also indicated by the presence of Vitrea sp. and Oxychilus hydatinus. Unfortunately, however, the sediments between 255 and 50cm contain little pollen and had to be bulked up to reach at least 100 per slide for the pollen graph. The resulting picture is thus somewhat sketchy, particularly in the absence of another reliable radiocarbon date from this part of the section. The poor pollen record may again be due to bad preservation and oxygenation, particularly since the land snail record is often comparatively high in this part of the core, when compared to the base. This is even more regrettable as the available documentary evidence for the

An increased sedimentation rate may be reflected by the drastic decrease of molluscs, with a concurrent decrease in the pollen sum in the sediments. The accumulation of nearly 200mm3 of macroscopic charcoal at 320cm may be significant in the light of this. Dated 530 – 660 AD (see above), it may perhaps indicate that the early Byzantine times in Malta were not free from strife, although of course it is not possible to differentiate from the charcoal whether a fire was started naturally, accidentally or intentionally. The pollen spectrum that possibly corresponds to Late Byzantine Malta may be at 305cm and is characterised by an overall drastic decrease in the pollen sum, with arboreal pollen accounting for less than 10% of the total sum. The absolute amounts of olive and particularly cereal pollen are lower than before, although percentagewise there appears to be little change (compare Figures 7.3.2. and 8.3.). Linum pollen occurs again, and may perhaps indicate flax retting. The amount of brackish water and marine molluscs decreased considerably (Figure 7.1.2.), perhaps because of the increased landderived sedimentation, as suggested by the increase in land snails and freshwater molluscs in the assemblages. The land snail species found between 300cm and 295cm indicate again a predominantly open country/karstland environment with steppe vegetation, possibly as a result of the increased erosion. It is very possible that at around 300cm the lagoon, with a calculated depth of only around 80cm, would have been too shallow to be used for shipping, which may have prompted the construction of warehouses on Jesuits Hill (see above, Chapter III). The Arab conquest of the late 9th century AD may possibly have occurred at 285cm. Charcoal macroremains from this level may perhaps be connected with the attack, which, according to some sources (Al-Bakri and Al-Himyari, see above Chapter III) are said to have resulted in the islands being abandoned and only visited by shipbuilders for wood supplies. The possible abandonment notion may perhaps be supported by a 112

Spanish Period onwards would have been valuable material to supplement the environmental evidence. Instead, the pollen and plant macro-remains add little to what is already known. Flax retting was an industrial activity apparently also under Spanish Rule (Wettinger, 1981: 15-16) and may have been carried out at Marsa as indicated by the presence of Linum pollen at around 155cm. According to the pollen spectra, the olive industry may have lost its former importance, but again, the paucity of pollen may provide an inaccurate picture. The introduction of citrus trees and cotton during the Middle Ages left no recognisable mark in the sediments, nor did the portrayed prosperity of the island in the Post Arab Period and the importance of the cumin exports (see Blouet, 1964).

environment from land to sea to land as a result of erosion, washovers and changing sea levels at the coring location of Marsa Core 1 resulted in great fluctuations with regards to the accumulation and preservation of the biological components. This taphonomic bias particularly affected pollen preservation, which makes an accurate interpretation of the vegetational changes based on pollen alone often difficult. In several instances, additional information on the possible vegetational changes could be inferred from the non-marine molluscan assemblages. Owing to erosion of hillslope sediments and their subsequent re-deposition in the Marsa Plain, an unknown amount of sediment components were recycled, which resulted at times in a reversed chronology as indicated by several radiocarbon dates. Hence, straightforward graphs and interpretations, where, for example, high amounts of charcoal in a previous forest environment lead to open country taxa with high amounts of cereal pollen and other anthropogenic indicators, thus fairly unequivocally indicating human impact on the environment (e.g. Patterson et al., 1988; Tinner et al., 1999), are absent from Marsa Core 1. Furthermore, the resolution of some of the data from Marsa Core 1 is low (e.g. pollen between 560cm to the top of the core), and thus the effects of, for example, several large local fires on the vegetation could only be inferred from changes in the molluscan assemblages.

Freshwater influence at Marsa becomes more pronounced from 300cm onwards, as indicated by the plant macroremains and pollen (see Appendix III and Figure 7.3.1.), but declines again when the sediments reach the present water table at 135cm. The sediments overlying the water table created marshland, in which it appears that predominantly Lactucaea (lettuce family) thrived, although the high presence of its pollen may be because of a taphonomic bias. Similarly, freshwater molluscs abound in significant quantities in this section of the core (see Figure 7.1.2.and Appendix II). The now extinct species Planorbis planorbis occurs for the last time at 220cm in Marsa Core 1 (around the late Post Arab Period/early Spanish Period, according to Figure 8.2.), and appears to be replaced eventually by Planorbis moquini, perhaps as a result of sandier substrata and a decreased vegetation in the stream (Giusti et al. 1995:184) The sandier substrata are perhaps due to increased sedimentation rates and erosion, as may be inferred from the increased presence of land snail burrowing species (Cecilioidae) as from 205cm onwards. The freshwater molluscs decrease significantly from 120cm upwards (see Figure 7.1.3. and Appendix), which may indicate a diminishing influence of the freshwater stream, probably due to the continuous silting up of the Marsa Plain. The majority of the land snails in this section are ubitquitous species, followed by open country species. Leaf litter associated species are very few and very irregularly spaced out along this part of the core. The influence of the sea, perhaps expressed in the presence of the odd brackish water mollusc between 95cm and 50cm may indicate that storm events could still trouble the increasingly drying marshland at Marsa up to at least the 15th century AD.

Several extreme deposition events are likely to have created hiatuses in the sedimentation rates. As a result, the radiocarbon dates provide important anchor points in the stratigraphy, but the ages of the sediments between these points may fluctuate considerably and should thus be regarded with caution. Despite all the above-mentioned difficulties, several important insights could be gained from the various analyses. The earliest scientific evidence from Marsa Core 1 for human-induced changes detected in the Marsa catchment and possibly in other areas of the Maltese Islands dates back to the Neolithic and consists of the sporadic occurrence of cereal pollen at 1040cm and 940cm. The impact of the introduction of this crop by the first settlers in the Neolithic may appear to have been insignificant at 1040cm – Pinus pollen actually increases and macroscopic amounts of charcoal are very low. The lack of pronounced variations in the land snail assemblages, which seem to suggest a predominantly open country/karstland environment with steppic vegetation from the base upwards, may perhaps support the suggested low impact.

8.5. Conclusion Human impact on the Mediterranean environment in the Holocene has been the subject of numerous studies (e.g. Carrión & Van Geel, 1999; Carcaillet et al., 2002; Grove & Rackham, 2002; Sadori & Narcisi, 2004) using often a combination of various indicators (e.g. pollen, charcoal, spores, molluscs, dinoflagellates). The application of the array of methods and techniques for an assessment of human-induced changes on the Maltese environment met several difficulties. For one, the continuously changing

By the next time cereal pollen occurs (at 940cm, dated at 4840-4610 cal. BC), also the pollen spectra indicate an open landscape. The ‘already open landscape’ from the Bronze Age (Trump, 1966) and, even earlier, the Zebbug Phase (Schembri & Hunt, forthcoming) can now perhaps be pushed back as far as the Ghar Dalam Phase of the Neolithic. How forested, if at all, the islands were in the early Holocene, however, is still debatable as the 113

16th century onwards possibly dwarfs any impact caused by earlier settlers. Small as the contribution and impact of the prehistoric settlers, Phoenician/Punic and Roman inhabitants may have been on any environmental degradation, not all of their activities should be judged prima facie unfavourably, as the introduction of cereal, olive and vine may, perhaps, also be considered comparatively beneficial.

evidence does not appear to be unequivocal. The absence of any forest or leaf litter associated land snails, of any herbaceous taxa associated with woodlands and the taphonomic bias associated with differential pollen preservation which favours the preservation of Pinus, particularly throw doubt on an extensively forested environment. Furthermore, Cupressaceae are low pollen producers with fragile grains and the presence of these trees may be indicative of steppe environment. Significant amounts of macroscopic charcoal that would indicate large scale burning of trees for agricultural land is also lacking.

This thesis concerned aspects of an inter-disciplinary approach to palaeoenvironmental/archaeological research, which, it is hoped, has not only added new knowledge about the archaeology of the Maltese Islands, but also to that of past climates and palaeoecology. It is suggested that this integrated approach should be applied for future excavations in the Maltese Islands and thus exploit the potential of environmental research. In this way, new data would be added and the picture of past human impacts could be refined. The results of the above research also show that a flexible methodological approach should be adapted as not all tested methodologies can be applied without taking the particulars of the specific context into considerations.

From the evidence of Marsa Core 1, it may be suggested that human impact on the environment may be mainly assessed by the introduction of plants for cultivation, which sometimes appears to have significantly changed the general plant spectra, particularly through weeds associated with cultivation that at times increase significantly. The introduction of walnut Juglans during the Phoenician/Punic phase was possibly short-lived and perhaps only possible because of favourable climatic conditions, maybe in combination with managed irrigation systems. The agricultural activity that may perhaps have led to the reduction of the perennial stream at Marsa may actually have contributed positively to the ground stability due to an increased vegetation cover, as suggested by the non-marine molluscan assemblages in this part of the core. Severe fire events left their traces in the sediments of Marsa Core 1 several times from the Bronze Age onwards, which possibly had short-term effects on the landscape, but in the long run the environment appears to have recovered.

It is hoped that suggestions and implications arising from the data presented here and its interpretation and implications will be considered widely and open debates that spark off further research.

Generally, the evidence discussed above does not offer much support for the theory that the prehistoric inhabitants caused an irreparable impact on the environment through large-scale slash-and-burn land clearances. Rather, it suggests that these people were responding to environmental changes that were possibly mainly caused by extreme weather events and climate and were making the best of the island’s resources in a struggle for survival. There is no unequivocal evidence that the massive erosion and deposition events shown in the stratigraphy of Marsa Core 1 have been caused by human action. Instead, circumstantial evidence suggests climate, weather and tectonics as the prime agents of change. Of course, depending on the exigencies of the times, exploitation of various resources may have taken place (e.g. tree felling for profit in the Middle Ages, but also olive cultivation in the Classical Period), but none of these actions may, so far, be linked to or solely blamed for, a degradation of the environment and/or the erosion that occurred since the Neolithic. The current pressure the Maltese Islands are subjected to, caused by industrial activity and a population too large to sustain, is unparalleled in the islands’ history. The Knights of St John, and later the British, have catapulted a rather backwater island over several centuries into the limelight of history. The ever-growing impact their presence and actions had on the environment from the 114

APPENDIX I Munsell soil colours of the moist and dried samples of Marsa Core 1

Depth (mtrs) 0 0,05

Sample No Marsa 1/1 1/2

0,1 0,15 0,2 0,25 0,3 0,35 0,4 0,45 0,5 0,55 0,6 0,65 0,7 0,75

1/3 1/4 1/5 1/6

1/7 1/8 1/9 1/10 1/11 1/12

0,8

1/13

0,85

1/14

0,9

1/15

0,95

1/16

1 1,05 1,1 1,15 1,2 1,25 1,3 1,35 1,4 1,45 1,5

1/17 1/18 1/19 1/20 1/21 1/22

1/23 1/24

1,55 1,6 1,65 1,7 1,75 1,8 1,85 1,9 1,95 2

Munsell Code

Description

Munsell 500µ

Description light yellowish brown

Munsell 63µ 10YR 7/3

10YR 5/3, 10YR 5/4 10YR 5/3, 10YR 5/4 10YR 5/3 10YR 6/4 10YR 6/4 10YR 6/4

brown, yellowish brown brown, yellowish brown brown light yellowish brown light yellowish brown light yellowish brown

10 YR 6/4

Description very pale brown

2.5Y 6/3

light yellowish brown

2.5Y 7/3

pale yellow

2.5Y 7/3 10 YR 6/4 10 YR 6/4 10YR 6/4

pale yellow light yellowish brown light yellowish brown light yellowish brown

2.5Y 7/3 2.5Y 7/4 2.5Y 7/4 10YR 7/4

pale yellow pale yellow pale yellow very pale brown

Compression loss

10YR 5/4 10YR 5/4 10YR 5/4 10YR 5/4 10YR 5/4 10YR 5/4, 10YR 5/6 10YR 5/4, 10YR 5/6 10YR 5/4, 10YR 5/6 10YR 5/4, 10YR 5/6 10YR 5/4, 10YR 5/6 10YR 5/3 10YR 5/3 10YR 5/3 10YR 5/3 10YR 6/3 10YR 6/3

yellowish brown yellowish brown yellowish brown yellowish brown yellowish brown yellowish brown, yellowish brown yellowish brown, yellowish brown yellowish brown, yellowish brown yellowish brown, yellowish brown yellowish brown, yellowish brown brown brown brown brown dull yellow orange dull yellow orange

10YR 7/3 2.5Y 6/4 10 YR 6/4 10 YR 6/4 10YR 6/4 10YR 6/4

very pale brown light yellowish brown light yellowish brown light yellowish brown light yellowish brown light yellowish brown

2.5Y 7/4 2.5Y 7/4 2.5Y 7/4 10YR 7/3 2.5Y 7/4 10YR 7/4

pale yellow pale yellow pale yellow very pale brown pale yellow very pale brown

10YR 6/4

light yellowish brown

2.5Y 7/4

pale yellow

10YR 6/4

light yellowish brown

10YR 7/4

very pale brown

10YR 6/4

light yellowish brown

2.5Y 7/4

pale yellow

2.5Y 7/4

pale yellow

10YR 7/3

very pale brown

10 YR 7/4 2.5Y 6/3 2.5Y 7/3 2.5Y 7/3 2.5Y 7/2 2.5Y 7/4

very pale brown light yellowish brown pale yellow pale yellow light grey pale yellow

10YR 7/4 2.5Y 7/4 2.5Y 7/4 2.5Y 7/4 2.5Y 7/4 10YR 7/3

very pale brown pale yellow pale yellow pale yellow pale yellow very pale brown

Compression loss brown brown, dull yellowish brown brown greyish yellow brown greyish yellow brown greyish yellow brown greyish yellow brown grey grey grey greyish red greyish red

10YR 7/3 10YR 7/3

very pale brown very pale brown

10YR 7/3 10YR 7/3

very pale brown very pale brown

1/25 1/26 1/27 1/28 1/29 1/30 1/31 1/32 1/33 1/34

10YR 5/3 10YR 5/3, 10YR 5/8 10YR 5/3 10YR 5/2 10YR 5/2 10YR 5/2 10YR 5/2 10YR 5/1 10YR 5/1 10YR 5/1 2.5YR 5/2 2.5YR 5/2

2.5Y 6/4 2.5Y 7/2 2.5Y 6/3 2.5Y 6/2 10 YR 7/2 10YR 7/2 2.5Y 6/2 2.5Y 6/2 10YR 7/2 2.5Y 6/4

light yellowish brown light grey light yellowish brown light brownish grey light grey light grey light brownish grey light brownish grey light grey light yellowish brown

10YR 7/3 2.5 7/4 2.5Y 7/4 2.5Y 7/3 10YR 7/2 10YR 7/2 2.5Y 7/3 2.5Y 7/3 10YR 7/2 10YR 6/4

2,05

1/35

10YR 6/3

pale brown

10YR 6/4

light yellowish brown

10YR 6/4

2,1 2,15 2,2

1/36 1/37 1/38

very pale brown pale yellow very pale brown

10YR 7/3 2.5Y 7/3 10YR 7/3

1/39 1/40 1/41

pale brown pale brown pale brown, reddish grey reddish grey reddish grey reddish grey

10 YR 7/3 2.5Y 7/3 10YR 7/3

2,25 2,3 2,35

10YR 6/3 10YR 6/3 10YR 6/3, 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

very pale brown pale yellow pale yellow pale yellow light grey light grey pale yellow pale yellow light grey light yellowish brown light yellowish brown very pale brown pale yellow very pale brown

2.5Y 6/2 2.5Y 6/1 2.5Y 6/1

light brownish grey grey grey

2.5Y 7/2 2.5Y 7/2 2.5Y 6/2

2,4 2,45 2,5 2,55 2,6 2,65 2,7 2,75 2,8 2,85

1/42 1/43 1/44 1/45 1/46 1/47 1/48 1/49 1/50 1/51

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey

2.5Y 6/1 2.5Y 6/1 2.5Y 6/1 2.5Y 6/1 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 5/2

grey grey grey grey light grey light grey light grey light grey light grey greyish brown

2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2

116

light grey light grey light brownish grey light grey light grey light grey light grey light grey light grey light grey light grey light grey light grey

Depth (mtrs) 2,9 2,95 3 3,05 0,31 3,15 3,2 3,25 3,3 3,35

Sample No 1/52 1/53 1/54 1/55 1/56 1/57 1/58 1/59 1/60 1/61

Munsell Code

Description

Munsell 500µ

Description

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey

5Y 5/2 5Y 5/2 2.5Y 6/2 5Y 6/2 5Y 7/2 1gley 6/10Y 1gley 6/10Y 1gley 7/10Y 5Y 6/1 5Y 6/1

olive grey olive grey light brownish grey light olive grey light grey greenish grey greenish grey light greenish grey grey grey

Munsell 63µ 5Y 7/2 5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 2.5Y 7/2 5Y 7/2 5Y 7/1 5Y 7/2 1gley 7/10Y

3,4 3,45 3,5 3,55 3,6 3,65 3,7 3,75 3,8 3,85 3,9 3,95 4 4,05 4,1 4,15 4,2 4,25 4,3 4,35 4,4 4,45 4,5 4,55 4,6

1/62 1/63 1/64 1/65

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey reddish grey reddish grey

1gley 6/10Y 5Y 5/1 1gley 5/10Y 1gley 5/10Y

greenish grey grey greenish grey greenish grey

1gley 6/10Y 5Y 7/1 1gley 6/10Y 1gley 5/10Y

1/66 1/67 1/68 1/69 1/70 1/71 1/72 1/73 1/74 1/75 1/76 1/77 1/78 1/79 1/80

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey

1/81 1/82 1/83 1/84 1/85 1/86 1/87 1/88 1/89 1/90 1/91 1/92 1/93 1/94 1/95 1/96 1/97 1/98 1/99 1/100

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey reddish grey

greenish grey grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey grey greenish grey greenish grey greenish grey greenish grey greenish grey / light greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey grey greenish grey

1gley 6/10Y 1gley 6/10Y 1gley 6/10Y 1gley 6/10Y 5Y 6/1 1gley 6/10Y 5Y 6/1 5Y 6/1 1gley 6/10Y 1gley 6/10Y 1gley 6/10Y 1gley 6/10Y 1gley 6/10Y 1gley 6/10Y 1gley 7/10

4,65 4,7 4,75 4,8 4,85 4,9 4,95 5 5,05 5,1 5,15 5,2 5,52 5,3 5,35 5,4 5,45 5,5 5,55 5,6

1gley 5/10Y 5Y 6/1 1gley 6/10Y 1gley 6/5GY 1gley 6/5GY 1gley 6/10Y 1gley 5/10Y 1gley 5/10Y 1gley 5/10Y 1gley 5/N 1gley 5/10Y 1gley 5/10Y 1gley 5/5GY 1gley 5/10Y 1gley 5/10Y// 7/10Y 1gley 5/10Y 1gley 5/10Y 1gley 5/10Y 1gley 5/10Y 1gley 5/10Y 1gley 6/10Y 1gley 5/10Y 1gley 6/10Y 1gley 6/10Y 1gley 6/5GY 1gley 5/10Y 1gley 5/10Y 1gley 5/10Y 1gley 5/10Y 1gley 5/10Y 1gley 5/10Y 1gley 6/10Y 1gley 5/10Y 5Y 5/1 1gley 5/10Y

1gley 6/10Y 1gley 6/10Y 2.5Y 6/1 5Y 6/1 5Y 6/1 5Y 7/1 5Y 6/1 5Y 6/1 1gley 6/10Y 5Y 6/1 5Y 6/1 5Y 6/2 5Y 6/1 5Y 6/1 1gley 6/10Y 1gley 6/10Y 1gley 6/10Y 1gley 6/10Y 5Y 6/2 1gley 7/10Y

5,65 5,7 5,75 5,8

1/101 1/102 1/103 1/104

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey reddish grey reddish grey

5Y 6/1 5Y 6/1 5Y 6/1 5Y 5/1

grey grey grey grey

2.5Y 6/1 2.5Y 6/1 2.5Y 7/1 2.5Y 6/2

5,85

1/105

reddish grey

5Y 5/1

grey

2.5Y 6/2

5,9 5,95 6 6,05 6,1

1/106 1/107 1/108

reddish grey reddish grey reddish grey

5Y 6/1 5Y 6/1 5Y 5/1

grey grey grey

2.5Y 7/1 2.5Y 6/1 2.5Y 7/1

Description light grey light grey light grey light grey light grey light grey light grey light grey light grey light greenish grey greenish grey light grey greenish grey greenish grey

Large boulder

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

Compression loss

117

greenish grey greenish grey greenish grey greenish grey grey greenish grey grey grey greenish grey greenish grey greenish grey greenish grey greenish grey greenish grey light greenish grey greenish grey greenish grey grey grey grey light grey grey grey greenish grey grey grey light olive grey grey grey greenish grey greenish grey greenish grey greenish grey light olive grey light greenish grey grey grey light grey light brownish grey light brownish grey light grey grey light grey

Depth (mtrs) 6,15 6,2 6,25 6,3 6,35 6,4 6,45 6,5 6,55 6,6 6,65 6,7 6,75 6,8 6,85 6,9

Sample No

Munsell Code

Description

Munsell 500µ

Description

Munsell 63µ

Description

1/109 1/110 1/111 1/112

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey reddish grey reddish grey

6,95 7 7,05 7,1 7,15

1/113 1/114 1/115 1/116 1/117

2.5 YR 5/1

reddish grey

5Y 5/1 5Y 6/2, 5Y 5/1 5Y 5/1 5Y 5/1, 10YR 8/4 5Y 6/1

grey light olive grey + grey grey grey, very pale brown

2.5Y 6/1 5Y 6/1 2.5Y 6/1 5Y 7/1

grey grey grey light grey

grey

2.5Y 6/1

grey

10 YR 4/1 10 YR 4/1

dark grey dark grey

5Y 6/1 5Y 5/1

grey grey

5Y 7/1 2.5Y 6/2

5Y 5/2 5Y 6/1, 10YR 8/4 5Y 5/1, 10YR 8/4 10YR 8/3

olive grey grey, very pale brown

10YR 7/2 10YR 7/2

light grey light brownish grey light grey light grey

7,2 7,25

1/118 1/119

10 YR 4/1 10 YR 4/1

dark grey dark grey

7,3

1/120

7,35

1/121

7,4 7,45 7,5 7,55 7,6 7,65 7,7 7,75 7,8 7,85 7,9 7,95 8 8,05 8,1 8,15 8,2 8,25 8,3 8,35

1/122 1/123 1/124 1/125 1/126 1/127 1/128 1/129 1/130 1/131 1/132 1/133 1/134 1/135 1/136 1/137 1/138 1/139 1/140 1/141

10 YR 4/1, 10 YR 5/2 10 YR 4/1, 10 YR 7/6 10 YR 7/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 10 YR 8/6 2.5 YR 6/1 10 YR 7/6 10 YR 7/6 10 YR 7/6 10 YR 7/6 10 YR 7/6 2.5 YR 5/2 2.5 YR 5/1

dark grey, greyish yellow brown dark grey, yellowish brown yellow yellow yellow yellow yellow yellow yellow yellow yellow yellow yellow yellow reddish grey yellow yellow yellow yellow yellow greyish red reddish grey

grey, very pale brown

10YR 7/2

light grey

very pale brown

10YR 7/3

very pale brown

10YR 8/3 10YR 8/3 10YR 8/3 10YR 8/3 10YR 8/3 10 YR 8/2 10 YR 8/2 10 YR 8/2 10 YR 8/2 10YR 8/3 10YR 7/3 10YR 8/3 2.5Y 7/2 10YR 8/3 10YR 7/4 10YR 7/4 10YR 7/4 10YR 7/4 2.5Y 6/1 2.5Y 7/1

very pale brown very pale brown very pale brown very pale brown very pale brown very pale brown very pale brown very pale brown very pale brown very pale brown pale yellow very pale brown light grey very pale brown pale yellow very pale brown very pale brown very pale brown grey light grey

10YR 7/3 10YR 7/3 10YR 7/3 10YR 7/4 10YR 8/3 2.5Y 7/4 2.5Y 7/3 2.5Y 7/3 2.5Y 7/4 10YR 8/2 10YR 7/3 2.5Y 7/4 2.5Y 7/2 10YR 7/4 10YR 7/4 10YR 7/4 10YR 7/4 10YR 7/4 2.5Y 6/3 2.5Y 6/2

reddish grey

2.5Y 6/1

grey

2.5Y 7/2

2.5 YR 5/1 2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey reddish grey

10YR 7/1 2.5Y 6/1 2.5Y 6/1

light grey grey grey

10YR 6/3 2.5Y 6/2 2.5Y 6/3

1/146

2.5 YR 5/1

reddish grey

2.5Y 6/1

grey

10YR 6/2

8,65

1/147

2.5 YR 5/1

reddish grey

2.5Y 6/2

light brownish grey

10YR 6/3

8,7 8,75

1/148 1/149

2.5 YR 5/1 2.5 YR 5/1

reddish grey reddish grey

2.5Y 5/1 10YR 7/1

grey light grey

2.5Y 6/3 10YR 6/3

8,8 8,85

1/150 1/151

10 YR 7/6 10 YR 7/6

yellow yellow

pale yellow, orange dull orange, light grey

2.5Y 7/3 2.5Y 7/3

8,9 8,95 9

1/152 1/153 1/154

10 YR 7/6 10 YR 7/6 7.5 YR 6/6

yellow yellow orange

2.5Y 6/4, 7/4 7.5YR 6/4, 10YR8/2 10YR 8/2 10YR 6/4 10YR 7/3, 6/4

very pale brown very pale brown very pale brown very pale brown very pale brown pale yellow pale brown pale brown pale yellow very pale brown very pale brown pale yellow light grey very pale brown very pale brown very pale brown very pale brown very pale brown very pale brown light yellowish brown light brownish grey light grey pale brown light brownish grey light yellowish brown light brownish grey pale brown light yellowish brown pale brown pale brown

8,4

1/142

2.5 YR 5/1

8,45 8,5 8,55

1/143 1/144 1/145

8,6

very pale brown light yellowish brown very pale brown

10YR 8/2 2.5Y 7/4 10YR 6/4

very pale brown pale yellow l. yellow. brown

Compression loss

Large boulder

Stone

118

Depth (mtrs)

Sample No

Munsell Code

Description

Munsell 500µ

Description

Munsell 63µ

Description

9,05 9,1 9,15 9,2 9,25

1/155 1/156 1/157 1/158 1/159

7.5 YR 6/6 10 YR 6/4 10 YR 6/4 7.5 YR 5/6 7.5 YR 5/6

orange light yellowish brown light yellowish brown bright brown bright brown

light yellowish brown pale yellow dull yellow orange bright brown very pale brown

10YR 7/3 2.5Y 7/3 2.5Y 7/3 10YR 7/4 2.5Y 7/4

very pale brown pale brown pale brown very pale brown pale yellow

9,3 9,35 9,4 9,45 9,5

1/160 1/161 1/162 1/163 1/164

7.5 YR 5/6 7.5 YR 5/6 7.5 YR 5/6 7.5 YR 7/6 7.5 YR 7/6

bright brown bright brown bright brown orange orange

10YR 7/3 10YR 8/3 2.5Y 7/3 10YR 7/4 10YR 7/4

very pale brown

9,55 9,6 9,65 9,7 9,75 9,8

1/165 1/166 1/167 1/168 1/169 1/170

7.5 YR 7/6 7.5 YR 7/6 7.5 YR 7/6 7.5 YR 7/6 7.5 YR 7/6 7.5 YR 7/6

orange orange orange orange orange orange

10YR 6/4 2.5Y 7/4 10YR 7/2 10YR 6/6 10YR 7/4, 7.5YR 6/4 10YR 7/4 10YR 7/4 10YR 6/4,7/2 10YR 7/4 7.5YR 6/4, 10YR8/3 10YR 7/4 10YR 7/4 10YR 7/4 10YR 7/4 10YR 7/3 7.5YR 5/6,8/4

9,85 9,9 9,95 10 10,05

1/171 1/172 1/173 1/174 1/175

2.5 YR 4/6 2.5 YR 4/6 2.5 YR 4/6 5 YR 5/6 5 YR 5/6

reddish brown reddish brown reddish brown bright reddish brown bright reddish brown

10,1 10,15 10,2

1/176 1/177 1/178

5 YR 5/6 2.5 YR 4/6 2.5 YR 4/6

bright reddish brown reddish brown reddish brown

10,25

1/179

2.5 YR 4/6

reddish brown

10,3

1/180

2.5 YR 4/6

reddish brown

10,35

1/181

2.5 YR 4/6

reddish brown

7.5YR 5/4

dull brown

10YR 6/4

10,4 10,45 10,5 10,55 10,6 10,65

1/182 1/183 1/184 1/185 1/186 1/187

2.5 YR 4/6 2.5 YR 4/6 2.5 YR 4/6 2.5 YR 5/6 2.5 YR 5/6 2.5 YR 5/6

reddish brown reddish brown reddish brown bright brown bright brown bright brown

7.5YR 6/6 7.5YR 5/4 7.5YR 6/4,4/4 10YR 6/4 10YR 6/4 10YR 6/4

orange dull brown dull orange, brown light yellowish brown light yellowish brown light yellowish brown

10YR 7/4 7.5YR 6/4 7.5YR 6/4 2.5Y 7/3 10YR 7/3 10YR 6/4

10,7 10,75 10,8

1/188 1/189 1/190

5 YR 4/4 5 YR 4/4 5 YR 4/4

dull reddish brown dull reddish brown dull reddish brown

7.5YR 4/4, 6/4 7.5YR 4/4 7.5YR 5/6,7/4

7.5YR 6/4 7.5YR 6/4 10YR 6/4

10,85 10,9

1/191 1/192

5 YR 4/4 5 YR 4/4

dull reddish brown dull reddish brown

10,95

1/193

5 YR 4/4

dull reddish brown

brown, dull orange brown bright brown, dull orange dull brown dull yellow orange, dull brown dull brown, light grey

11

1/194

5 YR 4/4

dull reddish brown

11,05

1/195

5 YR 4/4

dull reddish brown

11,1 11,15

1/196 1/197

5 YR 4/4 5 YR 4/4

dull reddish brown dull reddish brown

7.5YR 5/4 10YR 7/4, 7.5YR5/4 7.5YR 5/4, 10YR8/2 7.5YR 5/4, 6/4, dull brown, dull orange, 8/2 grey 10YR 7/4, 6/4 very pale brown, dull yellow orange 7.5YR 5/4, 8/3 dull brown, light orange 7.5YR 8/3, 6/4 light orange, dull yellow orange

very pale brown very pale brown light yellowish brown very pale brown dull orange, light yellow orange very pale brown very pale brown very pale brown very pale brown very pale brown bright brown, light yellow orange 10YR 6/4 very pale brown 7.5YR 6/6 orange 7.5YR 6/6 orange 10YR 7/4, 8/4 very pale brown 7.5YR 5/6, 7/4 bright brown, dull orange 7.5YR 5/6 bright brown 7.5YR 5/6 bright brown 7.5YR 5/6,7/4 bright brown, dull orange 5YR 5/6, bright reddish brown 7.5YR7/4 7.5YR 6/6 orange

119

10YR 7/4 10YR 7/4 10YR 7/3 10YR 7/4 10YR 7/4 10YR 6/4

pale brown very pale brown very pale brown

7.5YR 6/4 7.5YR 6/4 7.5YR 6/4

very pale brown very pale brown very pale brown very pale brown very pale brown light yellowish brown dull orange dull brown very pale brown very pale brown light yellowish brown dull orange dull orange dull orange

7.5YR 6/4

dull orange

10YR 6/4

10YR 7/3 10YR 7/3

light yellowish brown light yellowish brown very pale brown dull orange dull orange pale brown very pale brown light yellowish brown dull orange dull orange light yellowish brown very pale brown very pale brown

10YR 7/4

very pale brown

10YR 7/4

very pale brown

10YR 7/3

very pale brown

2.5Y 7/3 10YR 8/3

pale brown very pale brown

7.5YR 6/4 7.5YR 5/4 10YR 7/4 10YR 7/4 10YR 6/4

APPENDIX II Shells found in Marsa Core 1, standardised per 150g sample

120

Shells /Marsa Core 1 Sample depth (cm) Pomatias sulcatus (Draparnaud) Pseudamnicola moussonii (Calcara) Carychium cf. schlickumi (Strauch) Lymnaea truncatula (Müller) Planorbis planorbis (Linnaeus) Planorbis moquini (Requien) Bulinus cf. truncatus (Audouin) Ancylus fluviatilis (Müller) Truncatellina callicratis (Scacchi) Granopupa granum (Draparnaud) Pleurodiscus balmei (Potiez & Michaud) Vitrea spp. Oxychilus hydatinus (Rossmässler) Ferussacciidae Ceciliodes acicula (Müller) Hohenwartiana hohenwarti (Rossmässler) Rumina decollata (Linnaeus) Clausiliidae Muticaria sp. Papillifera papillaris (Müller) Xerotricha apina (Lamarck) Trochoidea spratti (Pfeiffer) Heliciidae Cernuella caruanae (Kobelt) Cochlicella acuta (Müller) Theba pisana (Müller) Cantareus apertus (Born) Pisidium casertanum (Poli) Gibbula adansonii (Payraudeau) Bittium reticulatum (da Costa) Cerithium spp. Pirenella conica (Blainville) Turitella communis (Risso) Alvania sp. Rissoa ventricosa (Desmarest) Hydrobia spp. Truncatella subcylindrica ( Linnaeus) Bolinus brandaris (Linnaeus) Hexaplex trunculus (Linnaeus) Nassarius sp. Cyclope neritea (Linnaeus) Columbella rustica (Linnaeus) Volvarina mitrella (Risso) Conus mediterraneaus (Hwass) Turbonilla lactea (Linnaeus) Retusa truncatula (Bruguière) Haminoea hydatis (Linnaeus) Cylichna cylindracea (Pennant) Ovatella myosotis (Draparnaud) Lithophaga lithophaga (Linnaeus) Pectinidae (Wilkes) Ostrea edulis (Linnaeus) Loripes lacteus (Linnaeus) Kellia suborbicularis (Montagu) Acanthocardia sp. Parvicardium exiguum (Gmelin) Cerastoderma sp. Tellina planata (Linnaeus) Solecurtus strigilatus (Linnaeus) Azorinus chamasolen (da Costa) Veneridae Tapes decussatus (Linnaeus) unknown

1/197 1120-15 1,47 -

1/196 1110 -

1/195 1105 -

1/194 1100 -

1/193 1095 1,34 -

1/192 1090 2,09 2,09 -

1/191 1085 -

1/190 1080 1,28 1,28 -

1/189 1075 1,07 -

1/188 1070 0,99 -

1/187 1065 1,37 2,74 -

1/186 1060 2,83 -

1/185 1055 2,52 -

1/184 1050 1,23 -

1/183 1045 -

-

2,03 -

-

-

1,34 -

2,09 -

1,83 -

1,28 1,28 1,28 -

1,07 1,07 1,07 1,07 4,28 -

0,99 0,99 0,99 1,98 0,99 -

2,74 2,74 2,74 1,37 5,48 1,37 -

2,83 2,83 8,49 2,83 -

2,52 2,52 2,52 2,52 2,52 2,52 2,52 -

1,23 1,23 1,23 -

0,97 ?0.97 -

121

Shells /Marsa Core 1 Sample depth (cm) Pomatias sulcatus (Draparnaud) Pseudamnicola moussonii (Calcara) Carychium cf. schlickumi (Strauch) Lymnaea truncatula (Müller) Planorbis planorbis (Linnaeus) Planorbis moquini (Requien) Bulinus cf. truncatus (Audouin) Ancylus fluviatilis (Müller) Truncatellina callicratis (Scacchi) Granopupa granum (Draparnaud) Pleurodiscus balmei (Potiez & Michaud) Vitrea spp. Oxychilus hydatinus (Rossmässler) Ferussacciidae Ceciliodes acicula (Müller) Hohenwartiana hohenwarti (Rossmässler) Rumina decollata (Linnaeus) Clausiliidae Muticaria sp. Papillifera papillaris (Müller) Xerotricha apina (Lamarck) Trochoidea spratti (Pfeiffer) Heliciidae Cernuella caruanae (Kobelt) Cochlicella acuta (Müller) Theba pisana (Müller) Cantareus apertus (Born) Pisidium casertanum (Poli)

1/182 1040 -

1/181 1035 -

1/180 1030 -

1/179 1025 -

1/178 1020 -

1/177 1015 -

1/176 1010 -

Gibbula adansonii (Payraudeau) Bittium reticulatum (da Costa) Cerithium spp. Pirenella conica (Blainville) Turitella communis (Risso) Alvania sp. Rissoa ventricosa (Desmarest) Hydrobia spp. Truncatella subcylindrica ( Linnaeus) Bolinus brandaris (Linnaeus) Hexaplex trunculus (Linnaeus) Nassarius sp. Cyclope neritea (Linnaeus) Columbella rustica (Linnaeus) Volvarina mitrella (Risso) Conus mediterraneaus (Hwass) Turbonilla lactea (Linnaeus) Retusa truncatula (Bruguière) Haminoea hydatis (Linnaeus) Cylichna cylindracea (Pennant) Ovatella myosotis (Draparnaud) Lithophaga lithophaga (Linnaeus) Pectinidae (Wilkes) Ostrea edulis (Linnaeus) Loripes lacteus (Linnaeus) Kellia suborbicularis (Montagu) Acanthocardia sp. Parvicardium exiguum (Gmelin) Cerastoderma sp. Tellina planata (Linnaeus) Solecurtus strigilatus (Linnaeus) Azorinus chamasolen (da Costa) Veneridae Tapes decussatus (Linnaeus) unknown

2,21 4,42 -

?1.1 -

1,23 1,23 1,23 1,23 3,69 -

1,03 1,03 1,03 1,03 -

1,1 -

1,17 1,17 -

122

1,13 -

1/175 1005 1,27 -

1/174 1000 1,29 -

1/173 995 0,89 -

1/172 990 -

1/171 985 1 -

1/170 980 0,71 -

1/169 975 2,61 -

1/168 970 0,48 0,48 -

1,13 1,13 1,13 1,13 -

-

1,29 -

-

1,15 -

1 1 -

0,71 0,71 1,42 -

1,31 0,65 1,96 0,65 0,65 1,96 0,65 -

1,93 0,97 4,35 0,48 0,48 2,42 1,45 -

1/167 965 1,12 0,56 0,56 -

1/166 960 0,51 1,03 -

1/165 955 0,56 0,56 0,56 -

1/164 950 1,2 1,2 1,2 -

1/163 945 -

1/162 940 1,61 4,86 3,23 -

1/161 935 0,81 0,81 0,81 -

1/160 930 0,78 -

1/159 925 -

1/158 920 -

1/157 915 0,64 1,27 0,64 0,64 -

1/156 910 0,91 0,91 0,91 -

1/155 905 -

1/154 900 0,91 -

1/153 895 2,62 3,93 -

1/152 890 9,26 9,26 -

1/151 885 1,08 1,08 -

1/150 880 5,82 -

1/149 875 -

1/148 870 0,83 0,83 -

1,43 1,43 10,01 5,72 -

1/146 860 2,41 2,41 4,82 9,64 9,64 -

3,35 0,56 0,56 0,56 1,12 0,56 0,56 2,79 0,56

0,51 1,54 0,51 0,51 3,59 0,51

1,68 0,56 0,56 0,56 0,56 0,56 2,79 0,56 -

1,2 1,2 1,2 3,6 1,2

5,32 5,32 15,96 -

12,1 1,61 1,61 12,1 0,81 0,81 2,42 1,61 1,61 10,48 2,42 2,42 0,81

4,03 2,42 0,81 3,23 0,81 0,81 0,81 1,61 4,03 2,42 -

0,78 0,78 0,78 0,78 1,56 1,56 -

0,82 -

1,1 1,1 3,3 1,1 4,4 1,1 -

3,18 0,64 0,64 0,64 2,55 0,64 0,64 0,64 0,64 0,64 3,82 0,64 0,64 -

4,55 0,91 0,91 0,91 0,91 0,91 0,91 1,82 -

5 1 1 1 1 6 2 1

1,82 0,91 0,91 0,91 1,82 0,91 -

7,86 1,31 1,31 1,31 1,31 1,31 1,31 1,31 7,86 1,31 ?1,31 3,93 -

23,15 9,26 4,63 4,63 4,63 4,63 32,41 9,26 4,63

2,16 3,24 1,08 1,08 1,08 2,16 2,16 7,56 3,24 -

3,88 3,88 1,94 1,94 5,82 1,94 1,94

5,46 10,92 5,46 5,46 5,46 21,84 46,38 -

13,33 5 1,67 1,67 1,67 13,33 0,83 0,83 7,5 0,83 0,83 10 0,83 0,83 5 0,83

1,43 81,51 14,3 1,43 4,29 10,01 2,86 44,33 4,29 11,44 8,58 1,43 4,29 5,72 15,73 2,86 62,92 14,3 1,43 2,86 2,86 2,86

120,5 14,46 9,64 7,23 60,25 2,41 7,23 16,87 9,64 2,41 14,46 4,82 69,89 16,87 ?2,41 12,05 4,82

123

1/147 865 1,43 2,86 -

Shells /Marsa Core 1 Sample depth (cm) Pomatias sulcatus (Draparnaud) Pseudamnicola moussonii (Calcara) Carychium cf. schlickumi (Strauch) Lymnaea truncatula (Müller) Planorbis planorbis (Linnaeus) Planorbis moquini (Requien) Bulinus cf. truncatus (Audouin) Ancylus fluviatilis (Müller) Truncatellina callicratis (Scacchi) Granopupa granum (Draparnaud) Pleurodiscus balmei (Potiez & Michaud) Vitrea spp. Oxychilus hydatinus (Rossmässler) Ferussacciidae Ceciliodes acicula (Müller) Hohenwartiana hohenwarti (Rossmässler) Rumina decollata (Linnaeus) Clausiliidae Muticaria sp. Papillifera papillaris (Müller) Xerotricha apina (Lamarck) Trochoidea spratti (Pfeiffer) Heliciidae Cernuella caruanae (Kobelt) Cochlicella acuta (Müller) Theba pisana (Müller) Cantareus apertus (Born) Pisidium casertanum (Poli)

1/145 855 1,62 1,62 4,86 8,1 -

1/144 850 2,1 1,05 2,1 3,15 -

1/143 845 -

1/141 835 1,98 1,98 1,98 1,98 5,94 -

1/140-2 830 0,77 0,77 0,77 2,3 10 -

1/140-1

1,65 3,3 4,95 1,65 -

1/142 840 2,53 5,06 -

Gibbula adansonii (Payraudeau) Bittium reticulatum (da Costa) Cerithium spp. Pirenella conica (Blainville) Turitella communis (Risso) Alvania sp. Rissoa ventricosa (Desmarest) Hydrobia spp. Truncatella subcylindrica ( Linnaeus) Bolinus brandaris (Linnaeus) Hexaplex trunculus (Linnaeus) Nassarius sp. Cyclope neritea (Linnaeus) Columbella rustica (Linnaeus) Volvarina mitrella (Risso) Conus mediterraneaus (Hwass) Turbonilla lactea (Linnaeus) Retusa truncatula (Bruguière) Haminoea hydatis (Linnaeus) Cylichna cylindracea (Pennant) Ovatella myosotis (Draparnaud) Lithophaga lithophaga (Linnaeus) Pectinidae (Wilkes) Ostrea edulis (Linnaeus) Loripes lacteus (Linnaeus) Kellia suborbicularis (Montagu) Acanthocardia sp. Parvicardium exiguum (Gmelin) Cerastoderma sp. Tellina planata (Linnaeus) Solecurtus strigilatus (Linnaeus) Azorinus chamasolen (da Costa) Veneridae Tapes decussatus (Linnaeus) unknown

34,02 6,48 6,48 16,2 32,4 1,62 3,24 6,48 1,62 1,62 6,48 6,48 30,78 4,86 1,62 3,24

21 5,25 1,05 1,05 16,8 1,05 2,1 5,24 ?1,05 1,05 23,1 1,05 2,1 3,15

36,3 8,25 19,8 3,3 6,6 1,65 8,25 1,65 21,45 4,95 8,25 -

53,13 15,18 5,04 2,53 5,06 5,06 2,53 2,53? 10,12 17,71 2,53 -

27,72 11,88 1,98 1,98 11,88 1,98 1,98? 3,96 1,98 1,98 9,9 1,98 5,94 -

7,7 1,54 2,3 6,15 1,54 1,54 2,3 11,54 1,54 2,3 1,54

124

1,2 1,2 1,2 1,2 2,4 10,8 -

1/139 825 0,86 2,59 4,31 0,86 -

1/138 820 1,71 11,11 -

1/137 815 1,63 0,81 0,81 0,81 1,63 -

1/136 810 0,79 -

1/135 805 3,06 1,02 -

1/134-2 800 0,64 1,27 1,27 0,64 0,64 7 7,64 -

1/133 795 1,72 8,62 -

1/132 790 1,22 2,44 6,1 -

25,2 9,6 2,4 1,2 4,8 4,8 3,6 1,2 2,4 2,4 1,2 12 3,6 2,4 -

-

-

-

-

8,16 1,02 1,02 2,04 2,04 1,02 1,02 1,02 2,04 12,24 4,08 2,04

155,43 8,92 0,64 7 6,37 3,18 42,04 1,91 3,18 3,18 1,91 9,55 1,91 8,28 1,27 5,73 2,55 36,3 7 0,64 1,27 3,82 13,38 -

-

-

1/131 785 2,12 2,12 -

1/130 780 0,97 0,97 -

1/129 775 1,25 2,5 -

1/128 770 1,42 4,26 -

1/127 765 1,55 1,55 -

1/126 760 1,72 -

1/125 755 1,43 -

1/124 750 5,5 2,75 8,25 -

1/123 745 3,94 -

1/122 740 -

1/121 735 1,48 1,48 -

1/120 730 1,57 1,57 4,71 4,71 -

1/119 725 2,04 2,04 3,06 5,1 1,02 -

1/118 720 1,02 1,02 1,02 5,1 1,02 4,08 5,1 -

1/117 715 5,04 25,2 2,52 -

1/116 710 5,36 10,72 5,36 -

1/114-5 700 4,89 4,89 24,45 24,45 -

1/113 695 1,33 1,33 1,33 -

1/112 690 1,39 1,39 1,39 -

1/111 685 1,08 1,08 2,16 4,32 -

-

-

1,25 -

1,42 1,42 -

1,55 -

-

1,43 -

-

-

-

1,48 -

1,57 1,57 1,57 3,14 1,57 4,71 1,57 1,57 14,13 3,14 3,14 -

1 2,04 -

1,02 1,02 2,04 2,04 1,02 4,08 -

2,52 2,52 2,52 2,52 5,04 -

5,36 -

19,56 4,89 4,89 53,79 4,89 9,78 4,89 4,89 -

26,6 5,32 5,32 2,66 22,61 1,33 3,99 2,66 11,97 1,33 2,66 -

22,24 4,17 2,78 11,12 34,75 1,39 1,39 1,39 2,78 5,56 27,8 4,17 1,39 1,39 -

20,52 3,24 1,08 51,84 1,08 1,08 2,16 2,16 27 6,48 1,08 -

125

Shells /Marsa Core 1 Sample depth (cm) Pomatias sulcatus (Draparnaud) Pseudamnicola moussonii (Calcara) Carychium cf. schlickumi (Strauch) Lymnaea truncatula (Müller) Planorbis planorbis (Linnaeus) Planorbis moquini (Requien) Bulinus cf. truncatus (Audouin) Ancylus fluviatilis (Müller) Truncatellina callicratis (Scacchi) Granopupa granum (Draparnaud) Pleurodiscus balmei (Potiez & Michaud) Vitrea spp. Oxychilus hydatinus (Rossmässler) Ferussacciidae Ceciliodes acicula (Müller) Hohenwartiana hohenwarti (Rossmässler) Rumina decollata (Linnaeus) Clausiliidae Muticaria sp. Papillifera papillaris (Müller) Xerotricha apina (Lamarck) Trochoidea spratti (Pfeiffer) Heliciidae Cernuella caruanae (Kobelt) Cochlicella acuta (Müller) Theba pisana (Müller) Cantareus apertus (Born) Pisidium casertanum (Poli)

1/110 680 1,16 1,16 3,48 2,32 2,32 -

1/109 675 1,39 1,39 2,78 1,39 -

1/108 600 2,34 1,17 1,17 -

1/107 595 0,67 0,67 0,67 0,67 4 0,67 2,67 -

1/106 590 -

1/105 585 0,99 0,99 0,99 -

1/104 580 5,61 2,8 0,93 -

1/103 575 1,05 -

1/102 570 3,12 -

1/101 565 1,06 1,06 1,06 1,06 1,06 -

1/100 560 3,66 1,22 7,32 2,44 -

1/099 555 1,22 3,66 1,22 9,76 6,1 -

1/098 550 4,24 1,06 1,06 1,06 2,12 4,24 4,24 -

1/097 545 2,1 6,3 6,3 -

Gibbula adansonii (Payraudeau) Bittium reticulatum (da Costa) Cerithium spp. Pirenella conica (Blainville) Turitella communis (Risso) Alvania sp. Rissoa ventricosa (Desmarest) Hydrobia spp. Truncatella subcylindrica ( Linnaeus) Bolinus brandaris (Linnaeus) Hexaplex trunculus (Linnaeus) Nassarius sp. Cyclope neritea (Linnaeus) Columbella rustica (Linnaeus) Volvarina mitrella (Risso) Conus mediterraneaus (Hwass) Turbonilla lactea (Linnaeus) Retusa truncatula (Bruguière) Haminoea hydatis (Linnaeus) Cylichna cylindracea (Pennant) Ovatella myosotis (Draparnaud) Lithophaga lithophaga (Linnaeus) Pectinidae (Wilkes) Ostrea edulis (Linnaeus) Loripes lacteus (Linnaeus) Kellia suborbicularis (Montagu) Acanthocardia sp. Parvicardium exiguum (Gmelin) Cerastoderma sp. Tellina planata (Linnaeus) Solecurtus strigilatus (Linnaeus) Azorinus chamasolen (da Costa) Veneridae Tapes decussatus (Linnaeus) unknown

17,4 3,48 1,16 2,32 92,8 2,32 2,32 2,32 1,16 1,16 8,12 81,2 8,12 2,32 1,16 1,16

30,58 4,17 1,39 11,12 8,34 104,25 1,39 6,95 1,39 1,39 5,56 2,78 93,13 15,29 1,39 1,39 4,17 -

46,8 8,19 1,17 2,34 5,85 33,93 2,34 2,34 1,17 38,61 2,34 1,17 1,17

35,33 4 1,33 2,67 2 67,3 2 0,67 1,33 0,67 2 0,67 3,33 2 1,33 28,67 4 0,67 0,67 -

7,2 3,6 48,6 1,8 12,6 1,8 ?1,8 -

2,97 0,99 54,46 1,98 0,99 24,75 3,96 -

20,56 3,74 7,48 9,35 4,67 101,87 0,93 5,61 1,96 0,93 54,57 5,61 0,93 0,93 -

11,55 1,05 2,1 7,35 3,15 40,95 1,05 2,1 18,9 4,2 1,05 -

33,28 3,12 1,04 3,12 1,04 4,16 125,84 -

25,44 6,36 13,78 12,72 5,3 226,84 1,06 12,72 1,06 1,06 2,12 3,18 71,02 12,72 1,06 -

78,08 25,62 47,58 4,88 10,98 315,98

95,16 18,3 8,54 4,88 20,74 392,84 1,22 2,44 12,2 1,22 4,88 1,22 1,22 10,98 86,62 31,72 1,22 3,66 6,1 2,44

98,58 16,96 18,02 4,24 14,84 138,86 1,06 4,24 6,36 4,24 2,12 5,3 66,78 11,66 1,06 1,06 -

105 8,4 16,8 12,6 189 16,8 2,1 4,2 2,1 4,2 2,1 84 10,5 2,1 -

126

2,08 2,08 2,08 62,4 3,12 1,04

7,32 1,22 2,44 2,44 4,88 2,44 84,18 15,86 4,88 10,98 -

1/096 540 4 1 2 1 1 1 1 12 6 1 -

1/095 535 6,54 1,09 1,09 1,09 1,09 1,09 2,18 1,09 2,18 5,45 1,09 -

1/094 530 2,32 2,32 1,16 2,32 1,16 -

1/093 525 1,07 1,07 1,07 1,07 -

1/092 520 -

1/091 515 -

1/090 510 2,36 -

1/089 505 1,54 1,54 -

1/088 500 -

1/087 495 -

1/086 490 -

1/085 485 2,14 1,07 1,07 1,07 -

1/084 480 1,11 1,11 1,11 -

1/083 475 2,42 6,05 2,42 2,42 6,05 1,21 6,05 4,84 -

1/082 470 4,5 1,5 3

1/081 465 -

1/080 460 1,14

1/079 455 1,16 2,32 -

3 274 17 17 2 20 1039 5 1? 10 23 1 6 9 18 10 215 44 3 4 -

3,27 94,83 18,53 15,26 16,35 337,9 6,54 1,09 4,36 7,63 4,36 4,36 10,9 5,45 94,83 21,8 1,09 3,27 -

1,16 25,52 12,76 8,12 9,28 390,92 11,6 5,8 1,16 6,96 1,16 2,32 2,32 3,48 47,56 4,64 3,48 1,16 -

2,14 12,84 5,35 1,07 5,35 150,87 2,14 1,07 3,21 2,14 1,07 1,07 1,07 22,47 4,28 1,07 -

16,05 8,56 2,14 2,14 203,3 1,07 5,35 19,26 6,42 -

1,1 5,5 2,2 5,5 1,1 248,6 5,5 1,1 2,2 2,2 1,1 1,1 7,7 5,5 -

8,26 25,96 14,16 2,36 7,08 464,92 3,54 1,18 7,08 3,54 1,18 2,36 23,6 8,26 -

4,62 6,16 16,94 12,32 4,62 3,08 514,36 1,54 1,54 3,08 1,54 1,54 3,08 1,54 7,7 10,78 1,54 -

12,9 18,06 9,03 9,03 2,58 1,29 368,94 1,29 1,29 2,58 3,87 1,29 1,29 6 7,74 -

14,24 21,36 8,9 3,56 3,56 208,26 1,78 5,34 8,9 1,78 16,02 7,12 1,78 -

5,64 11,28 5,64 22,56 169,2 5,64 5,64 22,56 5,64 -

1,07 28,89 1,07 4,28 169,06 3,21 2,14 1,07 1,07 1,07 2,14 36,38 11,77 2,14 -

1,11 4,44 1,11 1,11 69,93 1,11 1,11 1,11 1,11 1,11 1,11 18,87 3,33 -

2,42 25,41 9,68 2,42 3,63 227,48 1,21 2,42 1,21 1,21 1,21 25,41 2,42 1,21

4,5 30 10,5 4,5 4,5 351 1,5 7,5 1,5 3 1,5 1,5 1,5 28,5 10,5 1,5 -

3,63 16,94 13,31 2,42 2,42 2,42 267,41 1,21 1,21 2,42 1,21 2,42 12,1 1,21 1,21 -

3,42 22,8 3,42 2,28 5,7 174,42 1,14 2,28 2,28 1,14 9,12 6,84 -

29 11,6 3,48 8,12 359,6 3,48 1,16 6,96 1,16 13,92 1,16

127

Shells /Marsa Core 1 Sample depth (cm) Pomatias sulcatus (Draparnaud) Pseudamnicola moussonii (Calcara) Carychium cf. schlickumi (Strauch) Lymnaea truncatula (Müller) Planorbis planorbis (Linnaeus) Planorbis moquini (Requien) Bulinus cf. truncatus (Audouin) Ancylus fluviatilis (Müller) Truncatellina callicratis (Scacchi) Granopupa granum (Draparnaud) Pleurodiscus balmei (Potiez & Michaud) Vitrea spp. Oxychilus hydatinus (Rossmässler) Ferussacciidae Ceciliodes acicula (Müller) Hohenwartiana hohenwarti (Rossmässler) Rumina decollata (Linnaeus) Clausiliidae Muticaria sp. Papillifera papillaris (Müller) Xerotricha apina (Lamarck) Trochoidea spratti (Pfeiffer) Heliciidae Cernuella caruanae (Kobelt) Cochlicella acuta (Müller) Theba pisana (Müller) Cantareus apertus (Born) Pisidium casertanum (Poli)

1/077 445 1,17 2,34 1,17 -

1/076 440 -

1/075 435 2,26 -

1/074 430 0,95 3,8 -

1/073 425 1,05 1,05 3,15 2,1 1,05

1/072 420 0,98 0,98 0,98 0,98 1,96 3,92 1,96 -

1/071 415 1,16 1,16 1,16 -

1/070 410 2,46 -

1/069 405 1,41 -

1/068 400 1,71 1,71 -

1/067 395 1,95 -

1/066 390 -

1/065 355 1,09 -

Gibbula adansonii (Payraudeau) Bittium reticulatum (da Costa) Cerithium spp. Pirenella conica (Blainville) Turitella communis (Risso) Alvania sp. Rissoa ventricosa (Desmarest) Hydrobia spp. Truncatella subcylindrica ( Linnaeus) Bolinus brandaris (Linnaeus) Hexaplex trunculus (Linnaeus) Nassarius sp. Cyclope neritea (Linnaeus) Columbella rustica (Linnaeus) Volvarina mitrella (Risso) Conus mediterraneaus (Hwass) Turbonilla lactea (Linnaeus) Retusa truncatula (Bruguière) Haminoea hydatis (Linnaeus) Cylichna cylindracea (Pennant) Ovatella myosotis (Draparnaud) Lithophaga lithophaga (Linnaeus) Pectinidae (Wilkes) Ostrea edulis (Linnaeus) Loripes lacteus (Linnaeus) Kellia suborbicularis (Montagu) Acanthocardia sp. Parvicardium exiguum (Gmelin) Cerastoderma sp. Tellina planata (Linnaeus) Solecurtus strigilatus (Linnaeus) Azorinus chamasolen (da Costa) Veneridae Tapes decussatus (Linnaeus) unknown

2,34 22,23 8,19 2,34 2,34 5,85 186,03 1,17 3,51 1,17 2,34 24,57 12,87 1,17 -

2,86 8,82 1,96 1,96? 1,96 204,9 0,98 0,98 3,81 0,98 0,98 12,75 3,81 0,98 -

4,52 23,73 7,91 5,65 1,13 13,56 544,66 1,13 1,13 1,13 9,04 1,13 1,13 23,73 11,3 -

5,71 34,29 16,19 3,81 3,81 15,24 1329,52 2,85 6,67 0,95 0,95 14,29 0,95 0,95 1,9 3,81 34,29 6,67 0,95 1,9 0,95 -

18,9 44,1 26,25 37,8 2,1 16,8 1536,15 6,3 2,1 1,05 2,1 7,35 4,2 1,05 2,1 2,1 5,25 67,2 39,9 2,1 1,05 -

13,73 18,63 38,24 11,76 8,16 1508,82 0,98 2,94 14,71 0,98 1,96 6,86 3,92 1,96 2,94 50 50,98 1,96 0,98 -

9,28 18,56 19,72 20,88 10,44 803,88 1,16 1,16 3,48 1,16 6,96 3,48 1,16 1,16 23,2 41,76 1,16 1,16 1,16

7,38 8,61 6,15 7,38 1,23 2,46 356,7 1,23 4,92 1,23 1,23 7,38 9,84 1,23 -

9,87 11,28 9,87 2,82 467 1,41 4,23 2,82 4,23 15,51 19,74 -

6,84 17,1 20,52 1,71 3,42 232,56 3,42 18,81 5,13 -

1,95 7,8 7,8 7,8 5,85 140,4 1,95 5,85 15,6 11,7 -

14,76 36,9 7,38 7,38 7,38 88,56 7,38 7,38 29,52 -

4,36 27,25 21,8 16,35 1,09 1,09 6,54 773,9 1,09 1,09 1,09 10,9 4,36 -

128

1/064 350 -

1/063 345 -

1/062 340 -

1/061 335 1,16 -

1/060 330 1,13 1,13 -

1/059 325 4,32 5,4 2,16 -

1/058 320 8,61 1,23 4,92 12,3 2,46 -

1/057 315 5,55 2,22 6,66 -

1/056 310 1,34 1,34 2,68 2,68 -

1/055 305 1,92 1,92 3,84 -

1/054 300 1,67 -

1/053 295 10 1 1 5 6 1 -

1/052 290 35,34 1,72 0,86 2,58 2,58 2,58 2,58 11,21 0,86 32,76 1,72 -

1/051 285 2,16 1,08? 4,32 5,4 -

1/050 280 1,06 -

1/049 275 1,21 -

1/048 270 0,81 -

1/047 265 1,05 -

1/046 260 0,9 0,9 0,9 1,8 -

1/045 255 1,11 1,11 -

1/044 250 0,38 2,26 0,38 0,75 0,75 0,75 3,02 4,15 -

7,26 9,92 15,73 8,68 2,42 4,84 476,74 1,21 2,42 7,26 26,62 14,52 -

2,32 3,48 1,16 9,28 1,16 171,68 2,32 12,76 10,44 -

3,72 110,36 7,44 -

1,16 3,48 4,68 8,12 76,56 2,32 1,16 2,32 -

-

28,08 2,16 -

30,75 2,46 1,23 1,23 4,92 -

1,11 1,11 1,11 -

-

5,76 1,92 1,92 1,92 -

6,68 1,67 1,67 -

1 79 4 2 1 -

132,76 0,86 12,07 1,72 1,72 2,58 -

3,24 1,08 -

-

-

-

-

7,27 1,8 0,9 0,9 -

12,21 1,11 3,33 13,32 1,11 -

4,15 2,26 2,64 1,89 -

3,69 1,23 -

129

Shells /Marsa Core 1 Sample depth (cm) Pomatias sulcatus (Draparnaud) Pseudamnicola moussonii (Calcara) Carychium cf. schlickumi (Strauch) Lymnaea truncatula (Müller) Planorbis planorbis (Linnaeus) Planorbis moquini (Requien) Bulinus cf. truncatus (Audouin) Ancylus fluviatilis (Müller) Truncatellina callicratis (Scacchi) Granopupa granum (Draparnaud) Pleurodiscus balmei (Potiez & Michaud) Vitrea spp. Oxychilus hydatinus (Rossmässler) Ferussacciidae Ceciliodes acicula (Müller) Hohenwartiana hohenwarti (Rossmässler) Rumina decollata (Linnaeus) Clausiliidae Muticaria sp. Papillifera papillaris (Müller) Xerotricha apina (Lamarck) Trochoidea spratti (Pfeiffer) Heliciidae Cernuella caruanae (Kobelt) Cochlicella acuta (Müller) Theba pisana (Müller) Cantareus apertus (Born) Pisidium casertanum (Poli)

1/043 245 0,78 4,65 0,78 -

1/042 240 1,16 3,48 13,92 -

1/041 235 1,72 0,86 0,86 1,72 2,59 -

1/040 230 4,23 4,23 -

1/039 225 2,63 5,26 -

1/038 220 1,02 5,1 1,02 1,02 -

1/037 215 0,88 ?0,88 0,88 0,88 7,89 9 0,88 -

1/036 210 1,02 1,02 1,02 1,02 1,02 1,02 15,3 4,08 1,02 -

1/035 205 0,78 3,12 0,78 8,59 3,12 -

1/034 200 1,81 1,36 0,45 0,45 0,45 1,81 2,71 7,24 10,41 1,36 -

1/033 195 11,76 4,2 1,68 1,68 0,84 2,52 0,84 0,84 1,68 18,49 7,56 1,68

1/032 190 1,8 0,9 0,9 0,9 0,9 6,3 9,91 -

1/031 185 3,39 4,52 1,13 -

1/030 180 1,85 1,85 1,85 2,78 2,78 0,93 3,7 13,89 -

1/029 175 0,97 2,91 0,97 7,77 6,8 -

Gibbula adansonii (Payraudeau) Bittium reticulatum (da Costa) Cerithium spp. Pirenella conica (Blainville) Turitella communis (Risso) Alvania sp. Rissoa ventricosa (Desmarest) Hydrobia spp. Truncatella subcylindrica ( Linnaeus) Bolinus brandaris (Linnaeus) Hexaplex trunculus (Linnaeus) Nassarius sp. Cyclope neritea (Linnaeus) Columbella rustica (Linnaeus) Volvarina mitrella (Risso) Conus mediterraneaus (Hwass) Turbonilla lactea (Linnaeus) Retusa truncatula (Bruguière) Haminoea hydatis (Linnaeus) Cylichna cylindracea (Pennant) Ovatella myosotis (Draparnaud) Lithophaga lithophaga (Linnaeus) Pectinidae (Wilkes) Ostrea edulis (Linnaeus) Loripes lacteus (Linnaeus) Kellia suborbicularis (Montagu) Acanthocardia sp. Parvicardium exiguum (Gmelin) Cerastoderma sp. Tellina planata (Linnaeus) Solecurtus strigilatus (Linnaeus) Azorinus chamasolen (da Costa) Veneridae Tapes decussatus (Linnaeus) unknown

3,9 0,78 0,78 -

4,64 1,16 5,8 1,16 -

12,93 1,72 1,72 0,86 0,86 -

9,87 8,46 7,05 1,41 1,41 -

10,53 4,39 2,63 1,76 -

9,18 1,02 3,06 1,02 -

0,88 -

1,02 1,02 -

0,78 -

1,81 0,45 0,45 -

7,56 1,6 5,88 0,84 -

0,9 0,9 -

3,39 4,52 -

4,65 2,78 -

5,83 5,83 0,97 2,91 -

130

1/028 170 0,83 0,83 0,83 0,83 0,83 0,83 1,67 12,5 26,67 -

1/027 165 0,9 0,9 2,7 6,31 3,6 -

1/026 160 8,28 12,42 1,38 2,76 4,14 1,38 1,38 2,76 1,38 1,38 1,38 11,04 27,6 2,76

1/025 155 8,59 10,94 1,56 2,34 2,34 1,56 22,66 32,03 1,56

1/024 150 1,3 3,9 10,4 1,3 2,6 1,3 3,9 1,3 1,3 2,6 15,6 14,3 -

1/023 145 0,6 1,2 0,6 17,26 5,95 1,2 -

1/022 125 1,28 0,64 0,64 11,54 2,56 0,64 0,64 -

1/021 120 0,93 0,93 0,93 0,93 0,93 0,93 0,93 11,21 2,8 1,87 -

1/020 115 1,74 7,83 1,74 -

1/019 110 0,78 1,55 0,78 1,55 0,78 14,73 1,55 0,78 -

1/018 105 0,78 3,13 2,34 1,56 -

1/017 100 0,65 0,65 3,25 1,3 1,3 -

1/016 95 1,31 -

8,3 0,83 1,67 10,83 1,67 0,83 -

0,9 2,7 -

-

-

1,3? s

-

0,64? -

-

-

-

-

-

131

0,65 1,31 3,27 0,65 0,65 11,76 5,23 -

1/015 90 0,9 0,9 7,27 0,9 -

1/014 85 0,78 2,34 2,34 0,78 0,78 5,47 3,9 4,67 -

1/013 80 0,75 0,75 0,75 0,75 5,22 2,99 2,24 -

1/012 75 0,81 0,81 9,76 5,69 4,06 -

1/011 70 0,68 1,35 16,9 6,76 2,7 -

1/010 65 2,07 2,07 1,38 0,69 -

1/009 60 0,69 0,69 1,38 9,66 1,38 -

1/008 55 4,05 0,68 1,35 -

1,96 -

-

-

0,75 -

-

-

-

-

-

Shells /M arsa C ore 1 Sam ple depth (cm ) Pom atias sulcatus (D raparnaud) P seudam nicola m oussonii (C alcara) C arychium cf. schlickum i (Strauch) Lym naea truncatula (M üller) Planorbis planorbis (L innaeus) P lanorbis m oquini (R equien) B ulinus cf. truncatus (A udouin) A ncylus fluviatilis (M üller) Truncatellina callicratis (Scacchi) G ranopupa granum (D raparnaud) P leurodiscus balm ei (Potiez & M ichaud) Vitrea spp. O xychilus hydatinus (R ossm ässler) Ferussacciidae C eciliodes acicula (M üller) H ohenw artiana hohenw arti (R ossm ässler) R um ina decollata (L innaeus) C lausiliidae M uticaria sp. P apillifera papillaris (M üller) X erotricha apina (L am arck) Trochoidea spratti (Pfeiffer) H eliciidae C ernuella caruanae (K obelt) C ochlicella acuta (M üller) Theba pisana (M üller) C antareus apertus (B orn) P isidium casertanum (Poli) G ibbula adansonii (Payraudeau) B ittium reticulatum (da C osta) C erithium spp. P irenella conica (B lainville) Turitella com m unis (R isso) A lvania sp. R issoa ventricosa (D esm arest) H ydrobia spp. Truncatella subcylindrica ( L innaeus) B olinus brandaris (L innaeus) H exaplex trunculus (L innaeus) N assarius sp. C yclope neritea (L innaeus) C olum bella rustica (L innaeus) V olvarina m itrella (R isso) C onus m editerraneaus (H w ass) Turbonilla lactea (L innaeus) R etusa truncatula (B ruguière) H am inoea hydatis (L innaeus) C ylichna cylindracea (Pennant) O vatella m yosotis (D raparnaud) Lithophaga lithophaga (L innaeus) Pectinidae (W ilkes) O strea edulis (L innaeus) Loripes lacteus (L innaeus) K ellia suborbicularis (M ontagu) A canthocardia sp. Parvicardium exiguum (G m elin) C erastoderm a sp. Tellina planata (L innaeus) Solecurtus strigilatus (L innaeus) Azorinus cham asolen (da C osta) V eneridae Tapes decussatus (L innaeus) unknow n

1/007 50 0,63 1,89 1,26 -

1/006 25 0,77 0,77 3,08 -

1/005 20 0,61 0,61 3,68 1,23 -

1/004 15 0,57 1,7 1,7 -

1/003 10 0,68 -

1/002 5 0,99 0,99 1,98 3,96 -

1/001 0 2,04 0,68 -

0,63 -

-

-

1,7 -

0,68 2,03 0,68 -

2,97 0,99 2,97 38,61 4,95 -

1,36 4,08 0,68 0,68 1,36 0,68

132

APPENDIX III Plant macro-remains of Marsa Core 1, not standardised.

133

Depth (cm) PREDOMINANTLY WOODY TAXA Cupressaceae Ficus carica Olea europaea Rubus caesius Prunus sp. Vitis vinifera

5 15 180 210 215 235 240 245 250 285 290 295 300 320 325 2 1

1 1

1

PREDOMINANTLY HERBACEOUS TAXA Ajuga reptans Anagallis arvensis Anthemis arvensis Boraginaceae, cf. Nonnea Brassica nigra Chenopodium album Chondrilla juncea Coronopus squamatus Euphorbia exigua Euphorbia helioscopia Fumaria sp. Poaceae Hedysarum coronarium Lactuca sp. Linaria arvensis Oxalis corniculata Papaver dubium Reseda luteola Rumex pulcher Trifolium sp. Verbena officinalis

1

1

1 1

2 3

1

1

1

1

1

1 1

HYGRO-HYDROPHYTES Care x sp. Equisetum Medicago marina Nasturtium officinalis Polygonum persicaria Potamogeton sp. Ranunculus sg. Batrachium Ruppia sp. Scirpus cernuus Sphagnum sp. Zanichellia palustris unidentified

1 2

1

1

3 3 1

1

1

134

1

1

1

4

1

1

5

2

330 345 350 390 395 400 405 410 415 420 425 430 435 440 445 450 455 460 465 470 475 480 485

2

2

1

1

2

2

2

1

1

1

1

1

1

1 1

1

1 1

1

1 2

1 2

1 1

1

1

1

1 1

1 2

1 1 3 1

1 1

1

2

2

1

4

2

2 1

1

1

1

1

18

1 1

1

1

1

1

1

3

2

2

1

135

1

3

2

1

3

3

1

8

Depth (cm) PREDOMINANTLY WOODY TAXA Cupressaceae Ficus carica Olea europaea Rubus caesius Prunus sp. Vitis vinifera PREDOMINANTLY HERBACEOUS TAXA Ajuga reptans Anagallis arvensis Anthemis arvensis Boraginaceae, cf. Nonnea Brassica nigra Chenopodium album Chondrilla juncea Coronopus squamatus Euphorbia exigua Euphorbia helioscopia Fumaria sp. Poaceae Hedysarum coronarium Lactuca sp. Linaria arvensis Oxalis corniculata Papaver dubium Reseda luteola Rumex pulcher Trifolium sp. Verbena officinalis HYGRO-HYDROPHYTES Care x sp. Equisetum Medicago marina Nasturtium officinalis Polygonum persicaria Potamogeton sp. Ranunculus sg. Batrachium Ruppia sp. Scirpus cernuus Sphagnum sp. Zanichellia palustris

510 515 520 525 530 535 540 550 555 560 565 570 580 585

1

1

2

1

1 1

1

1 1

1

1

1

1 1 2

1

1 1

1

1

1

1 1

2

1

1

1 7

unidentified

136

1

1

1

76

1

3

4

2

4

2

1

1

2

1

1

595 600 675 680 685 690 695 700 720 775 785 800 825 830 835 840 845 850 855 860 865 875 915 1

1 1

1

1

1

2

1

1

1

1

1

1

1 1 1 1

1

3 1

1

1 2 2 13 2

1

1

1

1

1

1

1 1

4 1

1 1

1 1

2 1 1

1

1

1 1

1

1 1

1

2

1

1 1

1

1

1 1

1

3

2

6 4

2 1

1

1

1

3

1 1

1 2 1

1

1

12

137

1

1

6

1

3 1

1

1

1

2

4

7

5

2

2

Depth (cm) PREDOMINANTLY WOODY TAXA Cupressaceae Ficus carica Olea europaea Rubus caesius Prunus sp. Vitis vinifera PREDOMINANTLY HERBACEOUS TAXA Ajuga reptans Anagallis arvensis Anthemis arvensis Boraginaceae, cf. Nonnea Brassica nigra Chenopodium album Chondrilla juncea Coronopus squamatus Euphorbia exigua Euphorbia helioscopia Fumaria sp. Poaceae Hedysarum coronarium Lactuca sp. Linaria arvensis Oxalis corniculata Papaver dubium Reseda luteola Rumex pulcher Trifolium sp. Verbena officinalis

935 940 945 960 965 970 975 1025 1050 1055 1060 1090 1100 2

1

1

1

1

1 1

1

1

1

1 1

5

1

138

1

1

1 1

1

HYGRO-HYDROPHYTES Care x sp. Equisetum Medicago marina Nasturtium officinalis Polygonum persicaria Potamogeton sp. Ranunculus sg. Batrachium Ruppia sp. Scirpus cernuus Sphagnum sp. Zanichellia palustris unidentified

1 3

1

REFERENCES Item marked with an asterisk (*) below were only available to the author through secondary sources.

AMARI, M. Biblioteca Arabo-Sicula: versione italiana. 2 Vols. Turin-Rome, 1880-1881. AMMERMANN, A.J. The Acconia Survey: Neolithic Settlement and the Obsidian Trade. London: Institute of Archaeology, 1985.

ABELA, G. F. Della Descrittione di Malta, 1647. Facsimile edition. Malta: Midsea Books, 1984. ABELA, G. F.; CIANTAR, G. A. Malta Illustrata: ovvero descrizione di Malta isola del mare siciliano e adriatico, con le sue antichità, ed altre notizie; divisa in quattro libri. Malta: Mallia, 1772.

ANTONIOLI, F.; KERSHAW, S.; RUST, D.; VERRUBBI, V. Holocene sea-level change in Sicily and its implications for tectonic models: new data from the Taormina area, northeast Sicily. Marine Geology, 2003, Vol. 3293, pp. 1-19.

ABULAFIA, D. Henry Count of Malta and his Mediterranean activities: 1203-1230. In Luttrell, A. (ed.), Medieval Malta, Studies on Malta before the Knights, London: British School at Rome, 1975, pp. 104-125.

AQUILINA, J. Maltese-English Dictionary. Vols. I and II. Malta: Midsea Books Ltd, 1990. ARISTOTLE. Meteorologie. Paderborn: Schöningh, 1955.

ADAMS, A. L. *Observations on the fossiliferous caves of Malta. Journal of the Royal Dublin Society, Vol. 4(2), 1863, pp. 11-19: *On bones of fossil Chelonians from the ossiferous caves and fissures of Malta. Quarterly Journal of the Geological Society of London, 1866, Vol. 22, pp. 594595. Notes of a naturalist in the Nile Valley and Malta. Edinburgh: Edmonston and Douglas, 1870. *On remains of Mastodon and other Vertebrata remains of the Miocene beds of the Maltese Islands. Quarterly Journal of the Geological Society London, 1879, Vol. 35, pp. 517-531.

ASHBY, T. Roman Malta. Journal of Roman History, 1915, Vol. 5, pp. 23-80. ASHBY, T.; BRADLEY, R.N.; PEET, T.E.; TAGLIAFERRO, N. Excavations in 1908-11 in Various Megalithic Buildings in Malta and Gozo. British School at Rome, 1913, Vol VI, 126pp. ASHMORE, P.; BRAYSHAY, B.A.; EDWARDS, K.J.; GILBERTSON, D.D.; GRATTAN, J.P.; KENT, M.; PRATT, K.E.; WEAVER, R.E. Allochthonous and autochthonous mire deposits, slope instability and palaeoenvironmental investigations in the Borve Valley, Barra, Outer Hebrides, Scotland. The Holocene, 2000, Vol. 10 (1), pp. 97-108.

AGIUS-VADALA, M.; GANADO, A. The pre-Siege maps of Malta, 1536-1563. Annual report and accounts, 30th September 1986, of the Investment Finance Bank Ltd., Malta, 1986, pp. 1-17.

ATHAR, M.; MAHMOOD, M. Taxonomic perspective of plant species yielding vegetable oils used in cosmetics and skin care products. African Journal of Biotechnology, 2005, Vol. 4 (1), pp. 36-44.

AICHELE, D.; GOLTE-BECHTLE, M. Was blüht denn da? Wildwachsende Blütenpflanzen Mitteleuropas. Stuttgart: Frankh-Kosmos, 1993.

AURIEMMA, R.; MASTRONUZZI, G.; SANSÒ, P. Middle to Late Holocene relative sea-level changes recorded on the coast of Apulia (Italy). Géomorphologie: relief, processus, environnement, 2004, Vol. 1, pp. 19-34.

AITKEN, M.J. Science-based dating in Archaeology. England: Longman, 1990. ALLEN, J.R.M.; BRANDT, U.; BRAUER, A.; HUBBERTEN, H.-W.; HUNTLEY, B.; KELLER, J.; KRAML, M.; MACKENSEN, A.; MINGRAM, J.; NEGENDANK, J.F.W.; NOWACZYK, N.R.; OBERHÄNSLI, H.; WATTS, W.A.; WULF, S.; ZOLITSCHKA, B. Rapid environmental changes in southern Europe during the last glacial period. Nature, 1999, Vol. 400, pp. 740-743.

AURIEMMA, R.; MASTRONUZZI, G.; SANSÒ, P.; ZONGOLO, F. The harbour of the mansio ad speluncas (Brindisi, Italy): a key to the lecture of sea level changes in the past 3500 years? WIT Transactions on The Built Environment, 2004, Vol. 79, pp. 5-15.

ALPHA, T.R.; GALLOWAY, J.P.; TINSLEY, J.C. Karst Topography—A computer animation and paper model. U.S. Geological Survey Open-File Report 97-536A, 1997, pp.1-19.

AZZARO, R.; BARBANO, M.S. Seismogenic features of SE Sicily and scenario earthquakes for Catania. In Faccioli, E. and Pessina, V. (eds.) The Catania Project: earthquake damage scenarios for high risk area in the Mediterranean, pp. 9-13. Rome: CNR-GNDT, 1999.

ALVISI, F.; VIGLIOTTI, L. Magnetic signature of marine and lacustrine sediments from central Italy (PALICLAS Project), 1996. Memorie dell’ Istituto Italiano d’Idrobiologia, Vol. 55, pp. 285-302.

AZZARO, R.; BARBANO, M.S.; SIROVICH, L.; PETTENATI, F.; BOBBIO, M. Seismotectonic framework and scenario earthquakes. In Faccioli, E. and Pessina, V. (eds.) The Catania Project: earthquake

139

damage scenarios for high risk area in Mediterranean, pp. 7-8. Rome: CNR-GNDT, 1999.

the

BARRY, R.G.; CHORLEY, R.J. Atmosphere, Weather and Climate. Seventh edition. London: Routledge, 1998.

AZZOPARDI, A. A New Geography of the Maltese Islands. Malta: St. Aloysius’ College Publication, 1995.

BECKER, E. Malta Sotteranea. Strassburg: Heitz and Muendel, 1913

BALDACCHINO, A.E.; STEVENS, D. T. (eds.). IsSigar Maltin – l-uzu u l-importanza. Malta: Dipartiment ghall-Harsien ta’ l-Ambjent, 2000.

BEGG, S. Farming Edible Snails – lessons from Italy. Canberra: Rural Industries Research and Development Corporation, 2003.

BALDACCHINO, J.G. Appendix C. In Annual report on the working of the Museum Department during 19331934, pp. XXIII-XXVI. Malta: Government Printing Office, 1933-34. Appendix C. In Annual report on the working of the Museum Department during 1934-1935, pp. XVI-XXII. Malta: Government Printing Office, 1934-35. Appendix C. Natural History Section. In Annual report on the working of the Museum Department during 19351936, pp. XXVII-XXXV. Government of Malta, 193636. Appendix B. Natural History Section. In Annual report on the working of the Museum Department during 19361937, pp. XVI-XXIV. Government of Malta, 1936-37. Appendix B. Natural History Section. In Report on the work of the Museum Department for 1937-38, pp. VIXIV . Malta: Government Printing Office, 1937-38. Archaeology. In Annual Report on the working of the Museum Department for 1947-48, pp. I-III. Malta: Government Printing Office, 1947-48.

BEHRE, K.-E. The role of man in European vegetation history. In Huntley, B. and Webb, T. III (eds.), Vegetation History, pp. 633-672. Dordrecht: Kluwer Academic Publishers, 1988.

BANDY, M. Population and History in the Ancient Titicaca Basin. PhD Dissertation. Berkeley: University of California, 2001.

BETTIS, E. A. III. Soil Morphologic Properties and Weathering Zone Characteristics as Age Indicators in Holocene Alluvium in the Upper Midwest. In Holliday, Vance T. (ed.), Soils in Archaeology, pp. 119-144. Washington: Smithsonian Institution Press, 1992.

BELL, M.; WALKER, M.J.C. Late Quaternary Environmental Change. Physical and Human Perspectives. Second Edition. Harlow: Pearson Prentice Hall, 2005. BELLANTI, P.F. Studies in Maltese History. Valletta: Empire Press, 1924. BEN TIBA, B.; REILLE, M. Recherches pollenanalytiques dans les montagnes de Kroumirie (Tunisie septentrionale): premières résultats. Ecologia Mediterranea, 1982, Vol. VIII (4), pp. 75-86. BETA ANALYTIC INC. Calendar Calibration at Beta Analytic. Fact Sheet. Beta Analytic Inc: Miami, 2005.

BARBANO, M.S.; COSENTINO, M. Historical Earthquakes and Eruptions in the Sicilian Straits. In Neuberg, J. (ed.), Seismic Phenomena Associated with Volcanic Activity: Complimenting Seismology with other Monitoring Techniques for Forecasting Purposes. Conference Proceedings, p. 23. Leeds: Working Group of the European Seismological Commission, 2003.

BIRKS, H.J.B. Reconstructing Environmental Impacts of Fire from the Holocene Sedimentary Record. In Clark, J.S., Cachier, H. and Goldammer, J.G. (eds.), Sediment records of biomass burning and global change, pp. 295311. Berlin-Heidelberg: Springer, 1997.

BARBARO, A. Degli avanzi d’ alcuni antichissimi edifizi scoperti in Malta l’anno 1768. Dissertazione storico-critica. Monograph. Malta, 1794.

BLACKMAN, D. J. Archaeological evidence for sea level change. Zeitschrift für Geomorphologie, Neue Folge, 2005, Vol. 137, pp. 61-70.

BARCELO, P. Hannibal – Stratege und Staatsmann. Stuttgart: Klett-Cotta, 2004.

BLAGG, T.F.C.; BONANNO, A.; LUTTRELL, A.T. Excavations at Hal Millieri, Malta : a report on the 1977 campaign conducted on behalf of the National Museum of Malta and the University of Malta. Msida: Malta University Press, 1990.

BARICH, B.E.; GIARUDI, C. The late hunting societies of Jebel Gharbi, Libya: settlement and landscape. Proceedings of the 14th International UISPP Congress. Lièges, 2–8 September 2001, Colloque 15.1: ‘Hunters vs Pastoralists in the Sahara: Material Culture and Symbolic Aspects’, forthcoming.

BLAGG, T.F.C.; LUTTRELL, A.T. Notes on San Pawl Milqi. In Blagg, T.F.C; Bonanno, A.; Luttrell, A.T. (eds.) Excavations at Hal Millieri, Malta : a report on the 1977 campaign conducted on behalf of the National Museum of Malta and the University of Malta, pp. 147-152. Msida: Malta University Press, 1990

BARNES, R.S.K. The brackish-water fauna of northwestern Europe. Cambridge: Cambridge University Press, 1994.

140

BLOGGS, S. Jr. Principles of Sedimentology and Stratigraphy. Third edition. Upper Saddle River: Prentice Hall, 2001.

BORG, J. Descriptive Flora of the Maltese Islands, including the ferns and flowering plants. Reprint 1976. Koenigstein: Otto Koeltz Science Publishers, 1937.

BLOUET, B. The distribution of Marshland in Malta during the seventeenth century. Journal of Maltese Studies, 1964a, Vol. 2, pp. 198-204. The changing landscape of Malta during the rule of the crusading Order of St. John of Jerusalem. Unpublished PhD dissertation. University of Hull, 1964b. The Story of Malta. Second revised edition. London: Faber and Faber, 1972.

BORG, J. J. A Checklist to the Quaternary avifauna of the Maltese Islands – preliminary note. In Mifsud, A., Savona Ventura, C. (eds.) Facets of Maltese Prehistory 1999. Malta: Prehistoric Society of Malta, 1999, pp.7789. BORG, J. The Story of the Maltese Forest. Malta: Media Centre, 1995 BOSWELL, D. M. The Social Prestige of Residential Areas. In Sultana, R.G.; Baldacchino, G. (eds.), Maltese Society. A Sociological Inquiry, pp. 133-162. Malta: Mireva Publications, 1994.

BOELHOUWERS, J. C.; DE GRAAF, P. J.; SAMSODIEN, M. A. The influence of wildfire on soil properties and hydrological response at Devil’s Peak, Cape Town, South Africa. Zeitschrift für Geomorphologie, 1996, Suppl. Vol. 107, pp. 1-10.

BOSWORTH, V.; FELTON, A.; WETTER, L. Paleoclimatic reconstruction based upon a sediment core from the Kalya region, Lake Tanganyika, East Africa. Nyanza Project Annual Report 2004, University of Arizona, 2004.

BOJANOWSKI, A. Als Grönland noch grün war. Welt am Sonntag, 15/06/2003. BONANNO, A. 1976-77 L’habitat Maltese in Età Romana. ΚΩΚΑΛΟΣ, XXII-XXIII, 1976-77, Vol. I, pp. 385-395. Distribution of villas and some aspects of the Maltese economy in the Roman Period. Journal of the Faculty of Arts, 1977, Vol. VI, No.4, pp. 73-81. Quintinus and the location of the temple of Hercules at Marsaxlokk. Melita Historica VIII (3), 1982, pp. 190-240 Roman Artistic Heritage. Proceedings of History Week 1984, 1986a, pp.1-12. Malta: The Historical Society. A Socio-Economic Approach to Maltese Prehistory – the Temple Builders, in Malta, Studies of its Heritage and History, 1986b, pp.17-46. Malta: Mid-Med Bank Ltd. The archaeology of Gozo: from prehistoric to Arab times, in Charles Cini (ed.), Gozo: the Roots of an Island, Malta: Said International, 1990a. Malta’s Role in the Phoenician, Greek and Etruscan Trade in the Western Mediterranean. Melita Historica, 1990b, Vol. 10 (3), pp. 209-224. The Birgu peninsula in prehistoric and classical times. In Bugeja, L.; Buhagiar, M.; Fiorini, S. (eds.), Birgu: a Maltese maritime city, pp. 15-30. Msida: Malta University Services, 1993. Malta – ein archäologisches Paradies. Revised edition. Malta: MJ Publications, 2000. Malta. Phoenician, Punic and Roman. Malta: Midsea Books Ltd, 2005.

BOURKE, M.C.; THORPE, M. Rainfall-triggered slope failures in eastern Ireland. Irish Geography, 2005, Vol. 38, pp. 1-22. BOWEN-JONES, H.; DEWDNEY, J.C.; FISHER, W.B. Malta - Background for Development. Durham: Durham Colleges in the University of Durham, 1961 BOWEN-JONES, H.; BEELEY, B.W. Malta before 1800 AD. In Bowen-Jones, H.; Dewdney, J.N. and Fisher, W.B. (eds.), Malta – Background for Development. Durham: Durham Colleges in the University of Durham, 1961, pp. 103-114. BOWEN-JONES, H.; CHARLTON, W.A. Economy and Trade. In Bowen-Jones, H.; Dewdney, J.N. and Fisher, W.B. (eds.), Malta – Background for Development. Durham: Durham Colleges in the University of Durham, 1961, pp. 115-125. BOWERMAN, N. Coring for climate: Grain-size analyses of lake sediments, Kigoma Bay, Lake Tanganyika, Tanzania. Nyanza Project Annual Report 2003, pp. 39-42. University of Arizona, 2003. BRADFORD, E. The Great Siege, Malta 1565. London: Penguin Books, 1964.

BONANNO, A.; FRENDO, A.; VELLA, N. Excavations at Tas-Silg, Malta. A preliminary report on the 19961998 campaigns conducted by the Department of Classics and Archaeology of the University of Malta. Mediterranean Archaeology, 2000, Vol. 13, pp. 67-114.

BRIFFA, J.M. New light on Fr Magri’s exploration of the Hypogeum: notes from correspondence with the British Museum. Malta Archaeological Review, 200203, Vol. 6, pp.41-46.

BORG, A. Case Officer’s Report for Application 930/00. Floriana: Malta Environment and Planning Authority, 2001.

BRINCAT, J. M. Malta 870-1054, Al-Himyari’s Account. Malta: Said International, 1991.

141

BRITISH PETROLEUM EXPLORATION COMPANY. Geological maps, Malta and Gozo. Chessington: Ordnance Survey, 1957.

CAGIANO DE AZEVEDO, M. Medieval Buildings Excavated at Tas-Silg and San Pawl Milqi. In Luttrell, A. (ed.), Medieval Malta, Studies on Malta before the Knights, London: British School at Rome, 1975, pp. 8895.

BRL INGENIÈRIE. Diagnostic préalable pour une gestion halieutique dans la zone marine du Parc Naturel Régional de Camargue. Nîmes: Direction Régionale de l’Environnement Provence-Alpes-Côte d’Azur, 2004.

CALCUTT, A. The status of Malta in Roman Times. The Malta Year Book 1956, 1956, pp. 37-40. Life in Malta in Roman Times. The Malta Year Book 1957,1957, pp.41-47.

BRONK RAMSEY, C. Radiocarbon Calibration and Analysis of Stratigraphy: The OxCal Program Radiocarbon, 1995, Vol. 37 (2), pp. 425-430. Development of the Radiocarbon Program OxCal, Radiocarbon, 2001, Vol. 43 (2A), pp. 355-363. OxCal Program v3.10., 2005, http://www.rlaha.ox.ac.uk/ oxcal/oxcal.htm#author

CALANCHI, N.; DINELLI, E.; LUCCHINI, F.; MORDENTI, A. Chemostratigraphy of late Quaternary sediments from Lake Albano and central Adriatic Sea core (PALICLAS Project). Memorie dell’ Istituto Italiano di Idrobiologia, 1996, Vol 55, pp. 247-263.

BROUWER, W.; STÄHLIN, A. Handbuch der Samenkunde für Landwirtschaft, Gartenbau und Forstwirtschaft. Frankfurt: DLG-Verlags-GmbH, 1955.

CALLEJA-GERA, P. Ghar Mirdum, Malta. Bronze Age Cave Dwelling on Dingli Cliffs, 2001, www.shurdington.org/gharmirdum/index.html

BROWN, A.G. Alluvial geoarchaeology. Floodplain archaeology and environmental change. Cambridge: Cambridge University Press, 1997.

CAMPBELL, I.D. Quaternary pollen taphonomy: examples of differential redeposition and differential preservation. Palaeogeography, Palaeoclimatology, Palaeoecology, 1999, Vol. 149, pp. 245-256.

BROWN, T.S. Byzantine Malta: A discussion of the sources. In Luttrell, A. (ed.), Medieval Malta, Studies on Malta before the Knights, London: British School at Rome, 1975, pp. 71-87.

CANTI, M. A mixed approach to geoarchaeological analysis. In Barham, Anthony J. and Macphail, Richard I. (eds.), Archaeological Sediments and Soils: Analysis, Interpretation and Management, pp. 183-190. London: Institute of Archaeology, University College London, 2000.

BRUNO, B. L’arcipelago maltese in età romana e bizantina. Attività economiche e scambi al centro del Mediterraneo. Bari: Edipuglia, 2004.

CARCAILLET, C.; BOUVIER, M.; FRECHETTE, B.; LAROUCHE A.C.; RICHARD, P.J. Comparison of pollen-slide and sieving methods in lacustrine charcoal analyses for local and regional fire history. The Holocene, 2001, Vol. 11 (4), pp. 467-476.

BRUSASCO, P. Dal Levante al Mediterraneo Centrale: la prima fase fenicia a Tas Silg, Malta. Journal of Mediterranean Studies, 1993, Vol. 3 (1), pp.1-29. BUHAGIAR, M. Medieval Churches in Malta. In Luttrell, A. (ed.), Medieval Malta, Studies on Malta before the Knights, London: British School at Rome, 1975, pp. 163-180. Late Roman and Byzantine Catacombs and Related Burial Places in the Maltese Islands. Oxford: BAR International Series 302, 1986.

CARCAILLET, C.; ALMQUIST, H.; ASNONG, H.; BRADWHAW, R.H.W.; CARRION, J.S.; GAILLARD, M.-J.; GAJEWSKI, K.; HAAS, J.N.; HABERLE, S.G.; HADORN, P.; MÜLLER, S.D.; RICHARD, P.J.H.; RICHOOZ, I.; ROESCH, M.; SANCHEZ GONI, M.F.; VON STEDINGK, H.; STEVENSON, A.C.; TALON, B.; TARDY, C.; TINNER, W.; TRYTERUD, E.; WICK, L.; WILIS, K.J. Holocene biomass burning and global dynamics of the carbon cycle. Chemosphere. 2002, Vol. 49, pp. 845-863.

BUSUTTIL, J. Pirates in Malta. Melita Historica, 1971, Vol. V (4), pp. 308-10. CACHIA, C; MIFSUD, C.; SAMMUT, P.M. The Marine Shelled Mollusca of the Maltese Islands. Part I. Archaeogastropoda. Marsa, Malta: Grima Printing and Publishing, 1991. The Marine Mollusca of the Maltese Islands, Part 2: Neotaenioglossa. Leiden: Backhuys Publishers, 1996. The Marine Mollusca of the Maltese Islands, Part 3: Prosobranchia to Sub-class Pulmonata, order Basommatophora. Leiden: Backhuys Publishers, 2000. The Marine Molluscs of the Maltese Islands. Part IV. The classes Caudofoveata, Solenogastres, Bivalvia, Scaphopoda and Cephalopodia. Leiden: Backhuys Publishers, 2004.

CARMINATI, E.; DOGLIONI, C. Mediterranean tectonics. Encyclopaedia of Geology pp.135-146. Stuttgart: Elsevier, 2004. CARRIÓN, J.S.; VAN GEEL, B. Fine-resolution Upper Weichselian and Holocene palynological record from Navarrés (Valencia, Spain) and a discussion about factors of Mediterranean forest succession. Review of Palaeobotany and Palynology, 1999, Vol. 106, pp. 209236.

142

CARRIÓN, J. S.; SÀNCHEZ-GÓMEZ, P.; MOTA J. F.; YLL, R.; CHAÌN, C. Holocene vegetation dynamics, fire and grazing in the Sierra de Gádor, southern Spain. The Holocene, 2003, Vol. 13 (6), pp. 839-849.

CIS (CORPUS INSCRIPTIONUM SEMITICARUM) *Inscriptiones Phoenicias Continens, 1883, Part 1, Vol. 1. CLARK, D. Building Logistics. In Cilia, D. (ed.), Malta Before History, pp. 366-377. Malta: Miranda Publishers, 2004.

CARROLL, F. The Potential Use of Harbour Sediment in the Construction of Past Environmental History. Unpublished BSc Dissertation, University of Huddersfield, 2001.

CLARK, J. S.; CACHIER, H.; GOLDAMMER, J. G.; STOCKS, B. (eds.). Sediment records of biomass burning and global change. Berlin-Heidelberg: Springer, 1997.

CARUANA GATTO, A. Disappearance of Spondylus gaederopus L. and other species from Maltese Waters. The Mediterranean Naturalist, issue of April 1st, 1892.

COLLOCA, F.; CARDINALE, M.; BELLUSCIO, A.; ARDIZZIONE, G. Patterns of distribution and diversity of demersal assemblages in the Central Mediterranean Sea. Estuarine, Coastal and Shelf Science, 2003, Vol. 56, pp.469-480.

CASSAR, L.; TANTI, C. M.; ROLE, A. Coastal Area Management in the Maltese Islands Priority Actions Programme. Split: Regional Activity Centre, 2005. CASSAR PULLICINO, J. Malta in 1575: Social Aspects of an Apostolic Visit. Melita Historica, 1956, Vol. 2 (1), pp. 19-41.

COMBOURIEU NEBOUT, N.; TURON, J.L., ZAHN, R.; CAPOTONDI, L.; LONDEIX, L.; PAHNKE, K. Enhanced aridity and atmospheric high-pressure stability over the western Mediterranean during the North Atlantic cold events of the past 50 k.y. Geology, 2001, Vol. 30, pp. 863-866.

CATT, J.A.; STAINES, S.J. Petrography of sediments and buried soils. In Preece, R.C. and Bridgland, D.R. (eds.), Late Quaternary Environmental Change in Northwest Europe. Excavations at Holywell Coombe, Southeast England, pp. 69-85. London: Chapman and Hall, 1998.

CONSTANTINE, J. A.; PASTERNACK, G. B.; JOHNSON, M. L. Logging effects on sediment flux observed in a pollen-based record of overbank deposition. Earth Surface Processes and Landforms, 2005, Vol. 30, pp. 813-821.

*CHETCUTI, D.; BUHAGIAR, A.; SCHEMBRI, P.J.; VENTURA, F. The climate of the Maltese Islands. Msida: University of Malta, 1992.

COOK, M.E.; MORROW, H. Anthropogenic Sources of Cadmium in Canada. National Workshop on Cadmium Transport Into Plants, Canadian Network of Toxicology Centres, Ottawa, Ontario, Canada, June 20-21, 1995.

CHONDROGIANNI, C.; ARIZTEGUI, D.; ROLPH, T.; JUGGINS, S.; SHEMESH, A.; RIETTI-SHATI, M.; NIESSEN, F.; GUILIZZONI, P.; LAMI, A.; MCKENZIE, J.A.; OLDFIELD, F. Millenial to interannual climate variability in the Mediterranean during the Last Glacial Maximum. Quaternary International, 2004, Vol. 122, pp. 31-41.

COOKE, J.H. *Notes on the ‘Pleistocene Beds’ of Gozo. Geological Magazine (III), 1891,Vol. 8, pp. 348-355. *The Har Dalam Cavern, Malta and its fossiliferous contents. Proceedings of the Royal Society, 1893, Vol. 54, pp. 274-283. *Notes on the ‘Pleistocene beds’ of the Maltese Islands. Geological Magazine (IV), 1896, Vol. 3, pp. 201-210.

CHURCH, M. Sedimentary Analysis of Soil Samples. In Neighbour, T. (ed.), Excavation of a Bronze Age Cairn at Olcote, Breasdete, Near Calanais, Isle of Lewis. Scottish Archaeology Internet Reports, 2005, Vol. 13, pp. 47-50. Available on http://www.sair.org.uk/

COUDÉ-GAUSSEN, P.; ROGNON, P.; RAPP, A.; NIHLÉN, T. Dating of peridesert loess in Matmata, south Tunisia, by radiocarbon and thermoluminescence methods. Zeitschrift für Geomorphologie, 1987, N.F. 31(2), pp. 129-144.

CICERO, M. T. Reden gegen Verres II-VI, translated and edited by Gerhard Krueger. Stuttgart: Reclam, 2002

CUTAJAR, D. St. Paul’s sea-voyage. The Times of Malta, 13/3/2006.

CIL (CORPUS INSCRIPTIONUM LATINARUM) *Inscriptiones Bruttiorum, Lucaniae, Campaniae, Siciliae, Sardiniae Latinae. Vol. X, edited by Mommsen, T. Reprinted 1963. Berlin: Reimer, 1883.

CUTAJAR, N. Arabes et Normands à Malte. Dossiers d’Archéologie, 2001, Vol. 267, pp. 76-85.

CILIA, D. (ed.) Malta before History. Malta: Miranda Publishers, 2004.

DARBYSHIRE, I.; LAMB, H.; UMER, M. Forest clearance and regrowth in northern Ethiopia during the last 3000 years. The Holocene, 2003, Vol. 13 (4), pp. 537-546.

CILIA, J.L. A Carved Model of a Niche from Tarxien Temples Reconsidered, in The Oracle, 2000, Issue 1, pp.11-22.

143

DART, C.J.; BOSENCE, D.W.J.; MCCLAY, K.R. Stratigraphy and structure of the Maltese graben system. Journal of the Geological Society, 1993, Vol. 150 (6), pp. 1153-1166.

DICKENS, A. F.; GELINAS, Y.; HEDGES, J. I. Physical separation of combustion and rock sources of graphitic black carbon in sediments. Marine Chemistry, 2004, Vol. 92, pp. 215-223.

DAVIS, L. W. Weed Seeds of the Great Plains. A Handbook for Identification. Kansas: University Press of Kansas, 1993

DIODORUS SICULUS. Griechische Weltgeschichte, Vol. I – X. Translated by Gerhard Wirth. Stuttgart: Hiersemann, 1993.

DE BONO, E. Editorial. Malta Archaeological Review, 1998, Issue 1, p.2.

DOLOMIEU, D. G. De. Voyages aux îles de Lipari fait en 1781, ou Notices sur les îles Aeoliennes pour servir à l’histoire des volcans: suivi d’un memoire sur une espèce de volcan d’air, et d’un autre sur la température du climat de Malthe et sur la différence de la chaleur réelle et de la chaleur sensible. Paris: J.-M. Boursy, 1783.

DE GRAAFF, L.W.S. Coastal changes near Puttalam and the Mi Oya lagoonal delta, NW Sri Lanka. Zeitschrift für Geomorphologie, 1989, Suppl. Vol. 73, pp. 87-109.

DRAGO, A. Tide tables 1993, Grand Harbour. MRN Technical Report Series, 1993. Sea Level Measurements in Malta. In IOC Workshop Report No.176, MedGLOSS Workshop and Coordination Meeting for the Pilot Monitoring Network System of Systematic Sea Level Measurements in the Mediterranean and Black Seas, 2000,. Annex IVb, pp.8-11.

DEPARTMENT OF CIVIL AVIATION. Percentage frequencies of wind direction and speed within specified ranges. Period: 1958-1987. Drawing 6220/88. Malta: Department of Civil Aviation, 1988. DESPOTT, G. Excavations conducted at Ghar Dalam (Malta) in the summer of 1917. Journal of the Royal Anthropological Institute, 1918, Vol. 48, pp.214-221, pl. 17-19. Appendix C. Report of the Curator of the Natural History Section. In Report on the work of the Museum Department for 1923-24, pp. VIII-XIV. Malta: Government Printing Office, 1923-24. Appendix B. Report of the Curator of the Natural History Section. In Report on the work of the Museum Department for 1925-26, pp. VIII-XI. Malta: Government Printing Office, 1925-26. 1926-27 Appendix B. Report of the Curator of the Natural History Section. In Report on the work of the Museum Department for 1926-27, pp. XII-XIV. Malta: Government Printing Office, 1926-27. Appendix B. In Report on the work of the Museum Department for 1928-29, pp. VII-X. Malta: Government Printing Office, 1928-29. Appendix B. Natural History Section. In Report on the work of the Museum Department for 1929-30, pp. XIIIXIV. Malta: Government Printing Office, 1929-30. Malacological and Concological Notes. Extract from Bulletin of the Museum, Vol. 1, No. 11, pp. 3-6. Malta: Empire Press, Valletta, 1930.

DUCIE, Earl of. *Geological chart of the Island of Malta. Valletta: Goodenough, 1854. DUPONT, L., Pollen and Dinoflagellate Cysts of the Upper 50m of Site 958. In Firth, J.V. (ed.), Proceedings of the Ocean Drilling Program, Scientific Results, 1998, Vol. 159T, pp. 23-30. EVANS, J. D. Malta. London: Thames and Hudson, 1959. Prehistoric Antiquities of the Maltese Islands. London: Athlone Press, 1971. EVANS, J.G. The Environment of Early Man in the British Isles. London: Book Club Associates, 1957. An Introduction to Environmental Archaeology. New York: Cornell University Press, 1978. EVANS, J. G.; O’CONNOR, T. Environmental Archaeology, Principles and Methods. Stroud: Sutton Publishing, 1999. EVANS, M. E.; HELLER, F. Environmental Magnetism. Principles and Applications of Enviromagnetics. San Diego: Elsevier Science, 2003

DEWDNEY, J.C. Relief and Landforms. In BowenJones, H.; Dewdney, J.C. and Fisher, W.B. (eds), Malta – Background for Development, pp.34-42. Durham: Durham Colleges in the University of Durham, 1961a. Hydrology and Watersupply. In Bowen-Jones, H.; Dewdney, J.C. and Fisher, W.B. (eds), Malta – Background for Development, pp.43-47. Durham: Durham Colleges in the University of Durham, 1961b.

FAURÈ, G. Li storia ta Malta u Ghaudex bil gzejjer tahhom u il g'rajjiet li saru fihom, mictuba bil Malti minn Giovanni Faurè. Malta: Andolfo and Magro, 1913. FAUST, D.; ZIELHOFER, C.; BAENA ESCUDERO, R.; DIAZ DEL OLMO, F. High-resolution fluvial record of late Holocene geomorphic change in northern Tunisia: climatic or human impact? Quaternary Science Reviews, 2001, Vol. 23, pp-1757-1775.

DICKENS, A. F.; GELINAS, Y.; MASIELLO, C. A.; WAKEHAM, S.; HEDGES, J. I. Reburial of fossil organic carbon in marine sediments. Nature, 2004, Vol. 427, pp.336-339.

FEILDEN, H.W.; MAXWELL, E.C. Post-Pliocene beds in Gozo. Il Barth (Malta), 1874, Vol. 1(23), p. 468.

144

FENECH, K. Environmental Change and Human Activity at Tas-Silg (Malta) from the Bronze Age to the Punic Period, based on an analysis of environmental samples from Trench C5. Unpublished BA dissertation, University of Malta, 2001a. Report on the excavation of a Punic tomb: The molluscan remains. Malta Archaeological Review, 2001b, Vol. 5, pp. 20-22.

antiguos en el mediterráneo occidental. I seminario: El patrimonio arqueológico submarino y los puertos antiguos. Alicante, 2004, pp. 127-145. GANADO, A. The funeral of Angelo Emo in Malta in 1792, a pictoral record. Proceedings of History Week, 1993, pp. 151-180. GANDERT, O.-F. Preliminary Report on the Animal Bones. In Trump, D., Skorba. Appendix III. London: The Society of Antiquaries, 1966.

FERRING, C. R. Alluvial Pedology and Geoarchaeological Research in Holliday, Vance T. (ed.), Soils in Archaeology, pp. 1-40. Washington and London: Smithsonian Institution Press, 1992.

GARDNER, J. J.; WHITLOCK, C. Charcoal accumulation following a recent fire in the Cascade Range, northwestern USA, and its relevance for fire history studies. The Holocene, 2001, Vol. 11 (5), pp. 541549.

FINSINGER, W.; TINNER, W. Minimum count sums for charcoal-concentration estimates in pollen slides: accuracy and potential errors. The Holocene, 2005, Vol. 15(2), pp. 293-297.

GIRAUDI, C. The Upper Pleistocene to Holocene sediments on the Mediterranean island of Lampedusa (Italy). Journal of Quaternary Science, 2004, Vol. 19 (6), pp. 537-545.

FISCHER, K.; STEPHAN, B. Eine pleistozäne Avifauna aus der Ghar Dalam-Höhle, Malta. Zeitschrift für geologische Wissenschaften, 1974, Vol. 2(4), pp. 515523.

GIUSTI, F; MANGANELLI, G.; SCHEMBRI, P.J. The non-marine molluscs of the Maltese Islands. Monografie XV. Torino: Museo Regionale di Scienze Naturali, 1995.

FOUACHE, E.; DESRUELLES, S.; PAVLOPOULOS, K.; DALONGEVILLE, R.; COQUINOT, Y.; PEULVAST, J.-P.; POTDEVIN, J.-L. Using beachrocks as sea level indicators in the insular group of Mykonos, Delos and Rhenia (Cyclades, Greece). Zeitschrift für Geomorphologie, Neue Folge, 2005, Vol. 137, pp. 37-43.

GODWIN, H. Appendix I, Report on a pollen sample from a late Borg in-Nadur cistern at Tal-Mejtin, in Report on the working of the Museum Department for the year 1961, p. 8. Malta: Department of Information. Half-life of radiocarbon. Nature, 1962, Vol. 195, p. 944.

FRANICINI CORTI, E.; LANZA, B. XVIII Congresso della Società Italiana di Biogeografia: note conclusive sulla storia del popolamento animale e vegetale delle isole circumsiciliane. Lavori della Società Italiana di Biogeografia, 1972, Vol. 3, pp. 911-918.

GOLDBERG, E.D. Black carbon in the Environment: Properties and Distribution. New York: John Wiley and Sons, 1985.

FRANCIS, C.F.; THORNES, J.B. Runoff Hydrographs from Three Mediterranean Vegetation Cover Types. In Thornes, J.B. (ed.), Vegetation and Erosion, pp. 363-384. Chichester: John Wiley and Sons Ltd, 1990.

GOVERNMENT OF MALTA. Papers relative to the Harbour extension and the Government property in the French Creek. Plate II. Malta Harbour Extension. Malta: Government of Malta, 1870.

FRENDO, A. J. Some observations on the investigation of the Phoenicians/Canaanites in the ancient Mediterranean world. Journal of Mediterranean Studies, 1993, Vol. 3 (2), pp. 169-174.

GRASSO, M.; REUTHER, C.-D. The western margin of the Hyblean Plateau (SE Sicily): a neotectonic transform system on the SE Sicilian foreland. Annales Tectonicae, 1988, Vol. II, (2), pp. 107-120.

FUCHS, T. *Das Alter der Tertiärschichten von Malta. Vienna, 1875.

GRASSO, M.; PEDLEY, H. M.; REUTHER, C.-D. The geology of the Pelagian Islands and their structural setting related to the Pantelleria rift (central Mediterranean Sea). Centro, 1985, Vol. 1 (2), pp. 1-19.

GALEA, J. Topography of Malta in Abela’s Times. In Leopardi, E.R. (ed.), Gian Francesco Abela – Essays in his honour by members of the “Malta Historical Society” on the occasion of the Third Centenary of his death (1655-1955). Malta: Department of Information, 1961. GALE, S.J.; HOARE, P.G. Quaternary Sediments. London: Belhaven Press, 1991.

GRATTAN, J.; GILBERTSON, D. Energy Dispersive XRay Micro-Analysis and the Geochemistry of Soils and Sediments, in Leveau, P.; Trement, F.; Walsh, K.; Barker, G. (eds.) Environmental Reconstruction in Mediterranean Landscape Archaeology, pp. 81-88. Oxford: Oxbow Books, 1999.

GAMBIN, T. Islands of the Middle Sea: an archaeology of a coastline. In De Maria, L.; Turchetti, R. (eds.), Evolución paleoambiental de los puertos y fondeaderos

GRECH, C. F. A History of the Maltese Forest until 1800 AD. Unpublished PhD dissertation. University of Aberdeen, 2001. 145

GRECH, H. Tremor jolts Malta. The Times of Malta, 09/01/2006, 2006a. Ground water under threat. The Times of Malta, 31/01/2006, 2006b.

HEDGES, J. I.; HU, F. S.; DEVOL, A.H.; HARTNETT, H. E.; TSMAKIS, E., KEIL, R. G. Sedimentary organic matter preservation: a test for selective degradation under oxic conditions. American Journal of Science, 1999, Vol. 299, pp. 529-555.

GROVE, A.T.; RACKHAM, O. The Nature of Mediterranean Europe, an Ecological History. New Haven and London: Yale University Press, 2001.

HEDGES, J. I.; KEIL, R.G. Sedimentary organic matter preservation: an assessment and speculative hypothesis. Marine Chemistry, 1995, Vol. 49, pp. 81-115.

GUIDIBONI, E.; COMASTRI, A.; TRAINA, G. Catalogue of ancient earthquakes in the Mediterranean area up to the 10th century. Rome: Istituto Nazionale di Geofisica, 1994.

HELBAEK, H. Report on Carbonized Grain from AF5 (GhD. Phase). In Trump, D., Skorba. Appendix IV, p. 53. London: The Society of Antiquaries, 1966.

GUILLAUMIER, A. Bliet u Rhula Maltin. Vol. II. Malta: Valletta Publishing, 1987.

HIGHAM, T. Radiocarbon Web Info. www.c14dating.com, 1999

GULIA, G. *Repertorio di Storia Naturale. Malta, 185859.

HIGUERA, P. E.; SPRUGEL, D. G.; BRUBAKER, L. B. Reconstructing fire regimes with charcoal from smallhollow sediments: a calibration with tree-ring records of fire. The Holocene, 2005, Vol. 15 (2), pp. 238-251.

GUPTA, S.K.; POLACH, H.A. Radiocarbon Dating Practices at ANU. Handbook. Radiocarbon Laboratory, Research School of Pacific Studies. Canberra: ANU, 1985.

HILL, J.N. Broken K Pueblo: Prehistoric Social Organisation in the American Southwest. Anthropological Papers, No. 18. Tuscon: University of Arizona, 1970.

HALLETT, D., J.; MATHEWES, R.W.; WALKER, R.C. A 1000-year record of forest fire, drought and lake-level change in southeastern British Columbia, Cananda. The Holocene, 2003, Vol. 13 (5), pp. 751-761.

HÖLBL, G. Ägyptisches Kulturgut auf Malta und Gozo. Vienna: Verlag der österreichischen Akademie der Wissenschaften, 1989.

HANF, M. Ackerunkräuter Europas mit ihren Keimlingen und Samen. Munich: BLV Verlagsgesellschaft mbH, 1984.

HOLLOWAY, R. Archaeology of Ancient Sicily. London: Routledge, 1999.

HARLE, K.J. Late Quaternary vegetation and climate change in southeastern Australia: palynological evidence from marine core E55-6. Palaeogeography, Palaeoclimatology, Palaeoecology, 1997, Vol. 131, Issues 3-4, pp. 465-483.

HOMER. The Odyssey. Translated and revised by Rieu, D. London: Penguin, 1991.

HARTNETT, H. E.; KEIL, R.G.; HEDGES, J.I.; DEVOL, A.H. Influence of oxygen exposure time on organic carbon preservation in continental margin sediments. Nature, 1998, Vol. 391, pp. 572-574.

HOPKINS, J. A.; MCCARTHY, F. M.G. Postdepositional palynomorph degradation in Quaternary shelf sediments: a laboratory experiment studying the effects of progressive oxidation. Palynology, 2001, Vol. 26, pp. 167-184.

HOPE MONCRIEFF, A.R. Classical Reprinted 1994. London: Senate, 1907.

HASLAM, S.; BORG, J. The River Valleys of the Maltese Islands. Malta: Islands and Small States Institute of the Foundation for International Studies, 1998.

Mythology.

HOROWITZ, A. J. Spatial and temporal variations in suspended sediment and associated trace elements – requirements for sampling, data interpretation and the determination of annual mass transport. Archiv für Hydrobiologie. Advances in Limnology. 1996, Vol. 47, pp. 515-536.

HASLAM, S.M.; SELL, P.D.; WOLSELEY, P.A. A Flora of the Maltese Islands. Msida: Malta University Press, 1977.

HOUEL, J. Voyage pittoresque des Îles de Sicile, de Malte et de Lipari. Paris, 1787.

HAVINGA, A.J. Investigation into the differential corrosion susceptibility of pollen and spores. Pollen et Spores, 1964, Vol 6 (2), pp. 621-635. A 20-year Experimental Investigation into the Differential Corrosion Susceptibility of Pollen and Spores in Various Soil Types. Pollen et Spores, 1984, Vol. 26 (34), pp. 541-558.

HOUSE, M.R.; DUNHAM, K.C.; WIGGLESWORTH, J.C. Geology and Structure of the Maltese Islands, in Bowen-Jones, H.; Dewdney, J.C. and Fisher, W.B. (eds.), Malta - Background for Development, pp. 25-33. Durham: Durham Colleges in the University of Durham, 1961 146

HUGHES, Q. The Building of Malta, 1530-1795. London: Tiranti, 1956.

Association of Stratigraphic Palynologists Foundation, 1996.

HUGHES, K. J. Persistent features from a palaeolandscape: the ancient tracks of the Maltese Islands. The Geographical Journal, 1999, Vol. 165 (1), pp.62-79.

JARVIS, I.; MORETON, J.; GERARD, M. Chemostratigraphy of Madeira Abyssal Plain, MiocenePleistocene Turbidites, Site 950. In Weaver, P.P.E, Schmincke, H.-U., Firth, J.V. and Duffield, W. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, 1998, Vol. 157, pp. 535-558.

HUNT, C. O. Recent advances in pollen extraction techniques: a brief review. In Fieller, N.R.J., Gilbertson, D.D. and Ralph, N.G.A. (eds.) Palaeobiological Investigations. Research design, Methods and Data Analysis, 1985, pp. 181-7. Oxford: BAR International Series 266. Tafoni (pseudokarst features) in the Maltese Islands. Cave and Karst Science, 1996, Vol. 23 (2), pp. 57-62. Quaternary deposits in the Maltese Islands: a microcosm of environmental change in Mediterranean Islands. In GeoJournal, 1997, 41.2: pp. 101-109. Palynology. In Bonanno, A.; Frendo, A.and Vella N.C. (eds), Excavations at Tas-Silg, Malta. A preliminary report on the 1996-1998 campaigns conducted by the Department of Classics and Archaeology of the University of Malta. Mediterranean Archaeology, 2000, Vol. 13, pp. 111-114.

JEAN STANLEY, D.; HAIT, A. K. Deltas, radiocarbon dating, and measurements of sediment storage and subsidence. Geology, 2000, Vol. 28 (4), pp. 295-298. JONES, T.P.; CHALONER, W.G.; KUHLBUSCH, T.A.J. Proposed Bio-geological and Chemical Based Terminology for Fire-altered Plant Matter, in Clark, J.S.; Cachier, H. and Goldammer, J.G. (eds.), Sediment records of biomass burning and global change, pp. 9-22. Berlin-Heidelberg: Springer, 1997. JORDANOVA, N.; PETROVSKY, E.; KOVACHEVA, M.; JORDANOVA, D. Factors Determining Magnetic Enhancement of Burnt Clay from Archaeological Sites. Journal of Archaeological Science, 2001, Vol. 28, pp. 1137-1148.

HUNT, C. O., SCHEMBRI, P. J. Quaternary environments and biogeography of the Maltese Islands. In Mifsud, A., Savona Ventura, C. (eds.) Facets of Maltese Prehistory 1999. Malta: Prehistoric Society of Malta, 1999, pp.41-76.

JORDANOVA, N.; JORDANOVA, D.; VENEVA, L.; YOROVA, K.; PETROVSKY, E. Magnetic Response in Soils and Vegetation to Heavy Metal Pollution – A Case Study. Environmental Science and Technology, 2003, Vol. 37, pp. 4417-4424.

HUNTLEY, B.; WATTS, W.A.; ALLEN, J.R.M.; ZOLITSCHKA, B. Palaeoclimate, chronology and vegetation history of the Weichselian Lateglacial: comparative analysis of data from three cores at Lago Grande di Monticchio, southern Italy. Quaternary Science Reviews, 1999, Vol. 18, pp. 945-960.

KEENAN, D. J. Volcanic ash retrieved from the GRIP ice core is not from Thera. Geochemistry, Geophysics, Geosystems, 2003, Vol. 4 (11), pp. 1-8. KEITH , A. Neanderthal Man in Malta. The Journal of the Royal Anthropological Institute of Great Britain and Ireland, 1924, Vol. 54, pp. 251-260.

HUTTON, F.W. Sketch of the Physical Geology of the Maltese Islands. Geological Magazine, London, 1866, Vol. 3(22), pp. 145-151.

KELLETAT, D. Holocene Sea Level Curve for the Eastern Mediterranean from Multiple Indicators. Zeitschrift für Geomorphologie, Neue Folge, 2004, Vol. 137, pp. 1-9.

HYDE, H. P.T. The Geology of the Maltese Islands. With special reference to Water Supply and the possibilities of Oil. Valletta: A.C. Aquilina and Co, 1955.

KEPPIE, L. Understanding Roman Inscriptions. London: Routledge, 2001.

ISSEL, A. Note sur une caverne à ossements de l’île de Malte. Matériaux pour l’Histoire de l’Homme, 1866, Vol. 2, pp. 242-246.

KIMBALL, B. A. Effect of colloidal iron and aluminium on the transport and transformation of metals in rivers affected by mine drainage. Proceedings of the U.S. Geological Survey (USGS) Sediment Workshop, February 4-7, 1997.

JAHNS, S.; HULS, M; SARNTHEIN, M. Vegetation and Climate History of West Equatorial Africa Based on a Marine Pollen Record off Liberia (Site GIK 16776) Covering the Last 400,000 Years. Review of Palaeobotany and Palynology, 1998, Vol. 102 (3-4), pp. 277-288. JANUS, H. Unsere Schnecken und Muscheln. Stuttgart: Franck’sche Verlagsbuchhandlung, 1958.

KOHLER, T.A. Ceramic breakage Rate Simulation: Population Size and the Southeastern Chiefdom. Newsletter of Computer Archaeology, 1978, Vol.14, pp. 1-18.

JANSONIUS, J.; MCGREGOR, D.C. (eds.). Palynology. Principles and Applications. Vol. 2. Texas: American

KOUTSOUBAS, D.; ARVANITIDES, C.; DOUNAS, C., DRUMMOND, L. Community structure and dynamics of 147

the Molluscan Fauna in a Mediterranean lagoon (Gialova lagoon, SW Greece). Belgian Journal of Zoology, 2000, Vol. 130 (Supplement 1), pp. 131-138.

LUBELL, D. Prehistoric edible land snails in the circumMediterranean: the archaeological evidence. In Brugal, J.J. and Desse, J. (eds.), Petits Animaux et Sociétés Humaines. Du Complément Alimentaires Aux Ressources Utilitaires. XXIVe rencontres internationals d’archéologie et d’histoire d’Antibes, 2004, pp. 41-62.

LAMBECK, K.; ANTONIOLI, F.; PURCELL, A.; SILENZI, S. Sea-level change along the Italian coast for the past 10,000 yr. Quaternary Science Reviews, 2004, Vol. 23, pp. 1567-1598.

LUTTRELL, A. T. Approaches to Medieval Malta. In Luttrell, A.T. (ed.), Medieval Malta. Studies on Malta before the Knights, pp. 1-70. London: The British School at Rome, 1975.

LAMBECK, K.; BARD, E. Sea-level change along the French Mediterranean coast for the past 30000 years", Earth and Planetary Science Letters, 2000, Vol. 175, pp. 203-222.

LYNCH, J. A.; CLARK, J. S.; STOCKS, B. J. Charcoal production, dispersal, and deposition from the Fort Providence experimental fire: interpreting fire regimes from charcoal records in boreal forests. Canadian Journal of Forest Research, 2004, Vol. 34, pp.16421656.

LANFRANCO, E. Guida alle escursioni a Malta, Aprile 1984. Società Botanica Italiana, Sezione Siciliana, 1984. The Vegetation of the Maltese Islands, in Phipps, L. and Stainfield, J. Environmental themes in the Mediterranean: a web based case study of the Maltese Islands, 1999.Plymouth: SEED Publications, University of Plymouth. website:http://www.science.plym. ac.uk/ departments/geography/malta/default.htm

MACDONALD, G.M.; LARSEN, C.P.S.; SZEICZ, J.M.; MOSER, K.A. The reconstruction of boreal forest fire history from lake sediments: A comparison of charcoal, pollen, sedimentological, and geochemical indices. Quaternary Science Reviews, 1991, Vol.10 (1), pp.53-71.

LANG, D.M. Soils of Malta and Gozo, in Bowen-Jones, H. Dewdney, J.C. and Fisher, W.B. (eds.), Malta Background for Development, pp. 83-99. Durham: Durham Colleges in the University of Durham, 1961.

MACLANE, M. Sedimentology. New York: Oxford University Press, 1995.

LANGENSCHEIDT-LONGMANN. Dictionary of Contemporary English. Gütersloh: Longman Group, 1981.

MALEC, F.; STORCH, G. Zur Kenntnis der jungpleistozänen Wühlmaus Pitymys melitensis (Mammalia, Rodentia). Zeitschrift für Säugetierkunde, 1970, Vol. 35, pp.75-80.

LEBLANC, S. An addition to Naroll’s Suggested Floor Area and Settlement Population Relationship. American Antiquity, 1971, Vol. 36, pp. 210-211.

MALLIA-MILANES, V. (ed.) Hospitaller Malta, 15301798. Studies on Early Modern Malta and the Order of St. John of Jerusalem. Malta: Mireva Publications, 1993.

LEIGHTON, R. Sicily before History. An archaeological survey from the Palaeolithic to the Iron Age. London: Duckworth, 1999.

MALONE, C.; STODDART, S.; BONANNO, A.; GOUDER, T.; TRUMP, D. Mortuary ritual of the 4th Millenium BC in Malta: The Zebbug Period chambered tomb from the Brochtorff Circle at Xaghra (Gozo). Proceedings of the Prehistoric Society, 1995, Vol. 61: 303-345.

LE ROUX, G.; WEISS, D.; CHEBURKIN, A.; RAUSCH, N.; GRATTAN, J.; KOBER, B.; KRACHLER, M.; SHOTYK, W. Heavy metals, especially lead, deposition recorded in an ombrotrophic peat bog near Manchester, United Kingdom. Journal de Physique IV France, 2003, Vol. 107, pp. 739-742.

MALONE, C.; BONANNO, A.; GOUDER, T.; STODDART, S.; TRUMP, D. The death cults of prehistoric Malta. Scientific American, Special Edition, 2005, pp. 76-83.

LEWIS, H. Ancient Malta. A Study of Its Antiquities. Gerrards Cross, Bucks: Colin Smythe Ltd, 1977.

MANCINELLI, G.; ROSSI, L. The influence of allochthonous leaf detritus on the occurrence of crustacean detritivores in the soft bottom macrobenthos of the Po River Delta (northwestern Adriatic Sea. Estuarine, Coastal and Shelf Science, 2002, Vol. 54, pp. 849-861.

LIBBY, W.F. Radiocarbon Dating. Second Edition. Chicago: Chicago University Press, 1955. LIGHTFOOT, R.R. The Duckfoot Site, Vol. 2: Archaeology of the House and Household. Cortez: The Crow Canyon Archaeological Center, 1994. LOWE, J.J.; WALKER, Quaternary Environments. Longman, 1984.

MANGION, J.J. Report on Human Jaw Fragments from the Floor of the GhD. Hut (condensed). In Trump, D., Skorba. Appendix I. London: The Society of Antiquaries, 1966.

M.J.C. Reconstructing 2nd Edition. Harlow:

148

MANHEIM, F. Contaminated-Sediment Database Development and Assessment in Boston Harbor. USGS Fact Sheet 078-99. 1999

Report on the work of the Museum Department for the year 1961. Malta: Department of Information, 1961. Report on the working of the Museum Department for the year 1963. Malta: Department of Information, 1963. Report on the working of the Museum Department for the year 1964. Malta: Department of Information, 1964. Report on the working of the Museum Department for the year 1965. Malta: Department of Information, 1965.

MAR (MUSEUM ANNUAL REPORT) Annual Report, 1908-09. Malta: Government Printing Office, 1908-09. The Annual report of the Curator of the Valletta Museum for the Financial Year 1909-10. Malta: Government Printing Office, 1909-10. The Annual report of the Curator of the Valletta Museum for the Financial Year 1910-11. Malta: Government Printing Office, 1910-11. The Annual report of the Curator of the Valletta Museum for the Financial Year 1913-14. Malta: Government Printing Office, 1913-14. Report on the Museum for 1914-15. Malta: Government Printing Office, 1914-15. Report on the working of the Museum Department for the years 1915-16-17. Malta: Government Printing Office, 1915-17. Report on the working of the Museum Department for the years 1917-18-19. Malta: Government Printing Office, 1917-19. Report on the work of the Museum Department for 192324. Malta. Government Printing Office, 1923-24. Report on the work of the Museum Department for 192526. Malta. Government Printing Office, 1925-26. Report on the work of the Museum Department for 192627. Malta. Government Printing Office, 1926-27. Report on the work of the Museum Department for 192829. Malta. Government Printing Office, 1928-29. Report on the work of the Museum Department for 192930. Malta. Government Printing Office, 1929-30. Annual Report on the working of the Museum Department during 1933-34. Malta: Government Printing Office, 1933-34. Annual Report on the working of the Museum Department during 1934-35. Malta: Government Printing Office, 1934-35. Annual Report on the working of the Museum Department during 1935-36. Malta: Government Printing Office, 1935-36. Annual Report on the working of the Museum Department during 1936-37. Malta: Government Printing Office, 1936-37. Report on the work of the Museum Department for 193738. Malta: Government Printing Office, 1937-38. Annual Report of the working of the Museum Department during 1938-39. Malta: Government Printing Office, 1938-39. Report on the work of the Museum Department for 194647. Malta Government Gazette, 1946-47, Supplement No. LXII, pp. 1681-1690. Annual Report on the Working of the Museum Department, 1947-48. Malta: Government Printing Office, 1947-48. Report on the Working of the Museum Department, 195556. Malta: Government Printing Office, 1955-56. Report on the Working of the Museum Department, 195758. Malta: Government Printing Office, 1957-58. Report on the Working of the Museum Department, 195960. Malta: Government Printing Office, 1959-60.

MARTIN, A. C.; BARKLEY, W. D. Seed Identification Manual. Reprint of 1961 Edition. Caldwell: The Blackburn Press, 2000. MASTRONUZZI, G.; SANSÒ, P. Holocene coastal dune development and environmental changes in Apulia (southern Italy). Sedimentary Geology, 2001, Vol. 150 (12), pp. 139-152. MAYERHOFER, K. The Mystery of Hagar Qim. Malta: Union Print, 1996. MAYR, A. Die Insel Malta im Altertum. Munich: Verlagsbuchhandlung, 1909. MCCONNELL, B.E. San Cono- Piano Notaro – Grotta Zubbia Ceramics in Sicilian Prehistory. PhD Dissertation. Providence: Brown University, 1985. MCLUSKY, D. S.; ELLIOT, M. The Estuarine Ecosystem. Third Edition. Oxford: Oxford University Press, 2004. MEADOWS, M.E.; ASMAL, O. Chronology, sedimentology and geochemistry of sediments at Verlorenvlei (Western Cape Province, South Africa) as evidence of anthropogenically-induced land degradation. Zeitschrift für Geomorphologie, 1996, Suppl. Vol. 107, pp. 45-62. MEIRE, P. The impact of bird predation on marine and estuarine bivalve populations: a selective review of patterns and underlying causes. In Dame, R. (ed.) Bivalve Filter Feeders in Estuarine and Coastal Ecosystem Processes, pp. 197-244. Berlin-Heidelberg: Springer Verlag, 1992. MENZEL, W. Estuarine and Marine Bivalve Mollusc Culture. Boca Raton: CRC Press Inc, 1990. MEPA (MALTA ENVIRONMENT and PLANNING AUTHORITY) Draft Grand Harbour Local Plan. Floriana, 1997. State of the Environment Report (SoER). Floriana, 1998. Grand Harbour Local Plan. Approved Plan. Area Policies. Floriana, 2002a. State of the Environment Report. Floriana, 2002b. State of the Environment Report, 2005. Floriana, 2006. METCALFE, C. R. Report on the botanical determination of charcoal samples. In Trump, D., Skorba. Appendix V. London: The Society of Antiquaries, 1966.

149

MIFSUD, A. The Subterranean Sanctuary at Hal Saflieni. In Mifsud, A., Savona Ventura, C. (eds.), Facets of Maltese Prehistory 1999. Malta: Prehistoric Society of Malta, 1999, pp.149-168.

NICHOLS, G. J.; CRIPPS, J. A.; COLLINSON, M. E.; SCOTT, A. C. Experiments in waterlogging and sedimentology of charcoal: results and implications. Palaeogeography, Palaeoclimatology, Palaeoecology, 2000, Vol. 164, Issues 1-4, pp. 43-56.

MIFSUD, A.; MIFSUD, S. Dossier Malta – Evidence for the Magdalenian. Malta: Proprint Company Ltd, 1997.

NOTT, J. The tsunami hypothesis – comparisons of the field evidence against the effects, on the Western Australian coast, of some of the most powerful storms on Earth. Marine Geology, 2004, Vol. 208, pp. 1-12.

MITCHELL, P.K.; DEWDNEY, J.C. Climate. In BowenJones, H., Dewdney, J.C. and Fisher, W.B. (eds.), Malta Background for Development, pp. 48-82. Durham: Durham Colleges in the University of Durham, 1961.

NOTT, J.; BRYANT, E. Extreme Marine Inundations (Tsunamis?) of Coastal Western Australia. The Journal of Geology, 2003, Vol. 11, pp. 691-706.

MIZZI, P. The National Museums: A Historical Background. Heritage, 1982, Vol. 44, pp. 867-875.

NOTT, J.; THOMAS, M.; PRICE, D. Alluvial fans, landslides and late Quaternary climatic change in the wet tropics of northeast Queensland. Australian Journal of Earth Sciences, 2001, Vol. 48, pp. 875-882.

MÖRNER, N. Sea level changes and crustal movements with special aspects on the eastern Mediterranean. Zeitschrift für Geomorphologie, Neue Folge, 2005, Vol. 137, pp. 91-102.

NSO (NATIONAL STATISTICS OFFICE) Environment Statistics. Valletta: National Statistics Office, 2002.

MOORE, P.D.; WEBB, J.A.; COLLINSON, M.E. Pollen Analysis. Second Edition. Oxford: Blackwell Science, 1991.

OLDFIELD, F.; MAHER, B.A.; DONOGHUE, J.: PIERCE, J. Particle-size related, mineral magnetic source sediment linkages in the Rhode River catchment, Maryland, USA. Journal of the Geological Society, 1985, Vol. 142 (6), pp. 1035-1046.

MORHANGE, C. Relative sea-level changes in Marseille’s ancient harbors (France) during the Late Holocene. Zeitschrift für Geomorphologie, Neue Folge, 2005, Vol. 137, pp. 23-28

OROSIUS, P. Historiarum adversum paganos libri VII. Vienna: Carl Zangmeister, 1882.

MORONI, A.; AZZARO, R.; STUCCHI, M.; BARBANO, M.S.; ALBANI, P. Damage in Catania on the occasion of the 1693 earthquake. In Faccioli, E., Pessina, V. (eds.), The Catania Project: earthquake damage scenario for high risk areas in the Mediterranean. Rome: CNR-GNDT, 1999, pp. 119-125.

ORAU Datelist 28 Archaeometry, 1999, Vol. 41(2), pp. 421-431. OVID. *Fasti. Munich: Heimeran, 1960. PACE, A. The Artistic Legacy of Small Island Communities: the Case of the Maltese Islands (50002500 BC), in Pace, A. (ed.) Maltese Prehistoric Art 50002500 BC, pp. 1-12. Malta: Fondazzjoni Patrimonju Malti, 1996. The Building of Megalithic Malta. In Cilia, D. (ed.), Malta Before History, pp. 18-41. Malta, Miranda Publishers, 2004

MORTER, J. Capo Alfiere and the Middle Neolithic Period in Eastern Calabria, Southern Italy . PhD Dissertation. Department of Anthropology, University of Texas at Austin, 1992. MUNSELL SOIL COLOR CHART Revised edition. New Windsor, New York,: Munsell Color, 1994 MURRAY, J. The Maltese Islands with special reference to their geological structure. Scottish Geographical Magazine, 1890, Vol. 6, pp. 449-488.

PACE ABELA, M. Earthquakes in Malta - Storms and Tremors in 1343. Heritage, no date, Vol. 76, pp. 1517 – 1520.

N.A. Earthquake. The Malta Times, 21/8/1886. Malta’s Folk Lore. The Malta Standard, 17/10/1890.

PALOMBA, C.F.; TEDESCO, R. ClimaticEnvironmental Changes during the Late Quaternary in the Hyblean Plateau Coastal Zone (S-E Sicily). Global Changes during the Late Quaternary. Conference Proceedings 2001, pp. 50-51. Rome: Accademia Nazionale dei Lincei, 2001.

NAEVIUS, G. Cn. Naevii Belli Pvnici carminis qvae svpersvnt Leipzig: Teubner, 1964. NEHRING, B. Die Maltesischen Inseln. Tübinger Geographische Sudien, 1966, Vol. 19.

PATTERSON, W. A. III.; EDWARDS, K. J.; MAGUIRE, D.J. Microscopic charcoal as a fossil indicator of fire. Quaternary Science Reviews, 1987, Vol.6, pp. 3-23.

NICHOLS, G. Sedimentology and Stratigraphy. London: Blackwell Science, 1999.

150

PATTERSON, W. A. III; BACKMAN, A.E. Fire and disease history of forests. In Huntley, B. and Webb, T. III (eds.), Vegetation History, pp. 603-632. Dordrecht: Kluwer Academic Publishers, 1988.

Europe. Excavations at Holywell Coombe, South-east England. London: Chapman and Hall, 1998. PREECE. R. C. Mollusca. In Preece, R.C. and Bridgland, D.R. (eds.), Late Quaternary Environmental Change in North-west Europe. Excavations at Holywell Coombe, South-east England, pp. 158-212. London: Chapman and Hall, 1998.

PATTERSON, W. A. III; SASSAMAN, K. E. Indian Fires in the Prehistory of New England, in Nicholas, G. (ed.) Holocene Human Ecology in Northeastern North America, pp. 107-135. Plenum Publishing Corporation, 1988.

PREECE, R.C.; BRIDGLAND, D.R.; SHARP, M.J. Stratigraphical investigations. In Preece, R.C. and Bridgland, D.R. (eds.), Late Quaternary Environmental Change in North-west Europe. Excavations at Holywell Coombe, South-east England, pp. 33-68. London: Chapman and Hall, 1998.

PEDLEY, M.; CLARKE, M.H.; GALEA, P. Limestone Isles in a Crystal Sea, the Geology of the Maltese Islands. Malta: PEG Publications, 2002. PEET, T.E. Rough Stone Monuments and their Builders. London and New York: Harper and Brothers, 1912.

PTOLEMY, C. The Geography of Claudius Ptolemy. English Translation. Mineola: Dover Publications, 1991.

PERCIVAL, J. B.; LINDSAY, P. J. Measurement of Physical Properties of Sediments, in Mudroch, A.; Azcue, J.M. and Mudroch, P. (eds.), Manual of PhysicoChemical Analysis of Aquatic Sediments, pp. 7-46. Boca Raton: CRC Press, 1996.

PYATT, F.B.; GILMORE, G.; GRATTAN, J.P.; HUNT, C.O.; MCLAREN, S. An Imperial Legacy? An Exploration of the Environmental Impact of Ancient Metal Mining and Smelting in Southern Jordan. Journal of Archaeological Science, 2000, Vol. 27, pp. 771-778.

PERKINS, E.J. The Biology of Estuaries and Coastal Waters. London: Academic Press, 1974

PYNE, S.J.; GOLDAMMER, J.G. The culture of fire. An introduction to anthropogenic fire history. In Clark, J.S., Cachier, H.; Goldammer, J.G.; Stocks, B.J. (eds.) Sediment records of biomass burning and global change, pp. 71-114. Berlin-Heidelberg-New York: SpringerVerlag, 1997.

PERRY, A. Mediterranean Climate. In King, R.; Proudfoot, L.; Smith, B. (eds.) The Mediterranean. Environment and Society, pp. 30-44. London: Arnold, 1997.

RADKE, L. C. Catchment clearing impacts on estuaries. AUSGEO News, 2002, 65, pp.6-7.

PETIT-MAIRE, N. Climatic crises and man in the Mediterranean basin: the last 20,000 years. In Briand, F. (ed.), Human records of recent geological evolution in the Mediterranean Basin- historical and archaeological evidence, pp. 17-24. Monaco: CIESM, 2003.

RAMRATH, A.; SADORI, L.; NEGENDANK, J. F.W. Sediments from Lago di Mezzano, central Italy: a record of Lateglacial/Holocene climatic variations and anthropogenic impact. The Holocene, 2000, 10 (1), pp. 87-95.

PFLEGER, V. Weichtiere. Prague: Artia, 1984. PIKE, G. The Animal Bones from the Xemxija Tombs. In Evans J.D., Prehistoric Antiquities of the Maltese Islands: A Survey. Appendix, p. 239-41. London: Athlone Press, 1971.

REIMER, P. J.; BAILLIE, M.; BARD, E.; BAYLISS, A.; BECK, J.W., BERTRAND, C.J.H.; BLACKWELL, P.G.; BUCK, C.E.; BURR, G.S.; CUTLER, K.B.; DAMON, P.E.; EDWARDS, R.L.; FAIRBANKS, R.G.; FRIEDRICH, M.; GUILDERSON, T.P.; HOGG, A.G.; HUGHEN, K.A.; KROMER, B.; MCCORMAC, G.; MANNING, S.; BRONK RAMSEY, C.; REIMER, R.W.; REMMELE, S.; SOUTHON, J.R.; STUIVER, M.; TALAMO, S.; TAYLOR, F.W.; VAN DER PFLICHT, J.; WEYHENMEYER, C.E. IntCal 04, Terrestrial radiocarbon age calibration, 26-0 ka BP. Radiocarbon, 2004, Vol. 46, pp.1299-1304.

PIRAZZOLI, P. A. World Atlas of Holocene Sea-Level Changes. Elsevier Oceanography Series, 1990, Vol. 58. Sea-Level Changes. The Last 20 000 years. Chichester: John Wiley and Son, 1998. PITMAN, W.; RYAN, W. Sintflut. Ein Rätsel wird entschlüsselt. Bergisch Gladbach: Bastei Lübbe, 2001. PLANNING AUTHORITY. The Malta Structure Plan. Vol. I and Vol. II. Floriana, 1990.

RENFREW, C. Before Civilization. London: Penguin Books, 1973. Ancient Europe is older than we thought. National Geographic Magazine, 1977, Issue 152, pp. 615-623.

POKORNY, J. A Colour Guide to Familiar Trees. London: Octopus Books, 1973.

RENFREW, C.; BAHN, P. Archaeology. Theories, Methods and Practice. Second Edition. London: Thames and Hudson, 1995.

PREECE, R. C.; BRIDGLAND, D.R. (eds.). Late Quaternary Environmental Change in North-west

151

RENFREW, J. M. Cultivated plants from Tarxien Cemetery. Antiquity, 1972, Vol. 46, p.144-145.

in the Island. Oxford Journal of Archaeology, 1999, Vol. 23 (1), pp. 45-60. Land use in prehistoric Malta. A re-examination of the Maltese ‘cart-ruts’. Oxford Journal of Archaeology, 2004, Vo. 23 (1), pp. 45-60.

REUTHER, C.-D. Grabenrandtektonik: Die Maghlaq Störungszone auf Malta. Oberrheinische Geologische Abhandlungen, 1984a, Vol. 33, pp.67-82. Tectonics of the Maltese Islands. Centro, 1984b, Vol.1 (1), pp. 1-20. Extensional tectonics within the central Mediterranean segment of the Afro-European zone of convergence. Memorie della Società Geologica Italiana, 1987, Vol. 38, pp. 69-80.

SAGONA, C.; FARRUGIA, R.; BORDA, K.; ZAMMIT, M.-E.; STOEGER, H.; GALEA, M. Pottery. In Bonanno, A.; Frendo, A.; Vella N.C. (eds), Excavations at Tas-Silg, Malta. A preliminary report on the 1996-1998 campaigns conducted by the Department of Classics and Archaeology of the University of Malta. Mediterranean Archaeology, 2000, Vol. 13, pp. 84-97.

REUTHER, C.-D.; OBERMEYER, H.; REICHERTER, K.; REISS, S.; KAISER, A.; BUCHMANN, T.; ADAM, J.; LOHRMANN, J.; GRASSO, M. Neotektonik und aktive Krustenspannungen in Südost-Sizilien und ihre Beziehungen zur regionalen Tektonik im Zentralen Mittelmeer. Mitteilungen des GeologischPaläonologischen Instituts der Universität Hamburg, 2002, Vol. 86, pp. 1-24.

SAID-ZAMMIT, G. Population, Land Use and Settlement on Punic Malta. A Contextual Analysis of the Burial Evidence. Oxford: BAR International Series 682, 1997. The Phoenician and Punic Necropoles of Rabat, Malta. Melita Historica, 2001, Vol. 13, pp. 127-146. SALA, M.; CALVO, A. Response of Four Different Mediterranean Vegetation Types to Runoff and Erosion. In Thornes, J.B. (ed.), Vegetation and Erosion, pp. 347362. Chichester: John Wiley and Sons Ltd, 1990.

RIDOUT-SHARPE, J. The Mollusca. In E. Peltenberg (ed.), Excavations at Kissonerga-Mosphilia, 1979-1992, Lemba Archaeological Project, Cyprus, Vol. II.1B. Department of Archaeology, University of Edinburgh, Occasional Paper 19, 1998, pp. 224-229.

SANGIORGI, F.; DINELLI, E.; GABBIANELLI, G.; BRINKHUIS, H. Late Quaternary Climatic and Productivity Changes in the Sicily Channel (Central Mediterranean Sea). Geophysical Research Abstracts, 2003a, Vol. 5, Abstract 01713.

ROBINSON, C. Lago Grande di Monticchio, southern Italy: A long record of environmental change illustrated by sediment geochemistry. Chemical Geology, 1994, Vol. 118, pp. 235-254.

SANGIORGI, F.; CAPOTONDI, L.; COMBOURIEU NEBOUT, N.; VIGLIOTTI, L.; BRINKHUIS, H.; GIUNTA, S.; LOTTER, A.F.; MORIGI, C.; NEGRI, A. Holocene seasonal sea-surface temperature variations in the southern Adriatic Sea inferred from a multiproxy approach. Journal of Quaternary Science, 2003b, Vol. 18 (8), pp. 723-732.

ROSSI, L.; COSTANTINI, M.L. Mapping the introhabitat variation of leaf mass loss rate in a brackish Mediterranean lake. Marine Ecology Progress Series, 2000, Vol. 203, pp.145-159. ROSSIGNOL-STRICK, M.; PLANCHAIS, N. Climate patterns revealed by pollen and oxygen isotope records of Tyrrhenian sea core. Nature, 1989, Vol. 342, pp. 413416.

SARRE, R. The morphological significance of vegetation and relief on coastal foredune processes. Zeitschrift für Geomorphologie, 1989, Suppl. Vol. 73, pp. 17-31.

ROUSSEAU, R. M. Detection limit and estimate of uncertainty of analytical XRF results. The Rigaku Journal, 2001, Vol. 18 (2), pp. 33-48.

SCHEMBRI, P. J. The fauna of the Maltese Islands: a review and analysis. In Ellul-Micallef, R. and Fiorini, S. (eds.) Collegium Melitense Quartercentenary Celebrations (1592-1992); Collected papers contributed by members of the academic staff of the University of Malta, pp. 541-573. Msida, Malta: University of Malta, 1992. Physical Geography and Ecology of the Maltese Islands: A Brief Overview. Options Méditerranéennes, 1993, Sér. B, Vol.7, pp.27-39. The Maltese Islands: climate, vegetation and landscape. GeoJournal, 1997, Vol. 41 (2), pp. 1-11.

RML (ROYAL MALTA LIBRARY). *Section 9, Fragment 108. Manuscript, no date. SADORI, L.; NARCISI, B. The Postglacial record of environmental history from Lago di Pergusa, Sicily. The Holocene, 2001, Vol. 11 (6), pp. 655-670. SADORI, L.; GIRAUDI, C.; PETITTI, P.; RAMRATH, A. Human impact at Lago di Mezzano (central Italy) during the Bronze Age: a multidisciplinary approach. Quaternary International, 2004, Vol. 113, pp. 5-17.

SCHEMBRI, P. J.; FALZON, A.; FENECH, K.; SANT, M. J. The Molluscan Remains. In Bonanno, A.; Frendo, A.; Vella N.C. (eds), Excavations at Tas-Silg, Malta. A preliminary report on the 1996-1998 campaigns conducted by the Department of Classics and

SAGONA, C. Silo or Vat? Observations on the Ancient Textile Industry in Malta and Early Phoenician Interests

152

Archaeology of the University of Malta. Mediterranean Archaeology, 2000, Vol. 13, pp. 102-109.

School at Rome and Accordia Research Centre, University of London, 1994.

SCHEMBRI, P.J.; HUNT, C.O. Terrestrial molluscs and the changing environment of the Xaghra Circle, Gozo, Maltese Islands. Forthcoming.

SMETTAN, H. W. Pollenanalytische Voruntersuchungen im Umfeld der bandkeramischen Siedlung. Bericht der römisch-germanischen Kommission. Vol. 79, pp.59-63. Mainz: Philip von Zabern, 1998.

SCHEMBRI, P.J.; LANFRANCO, E. Development and the natural environment in the Maltese Islands. In Lockhart, D.G.; Draka-kis-Smith, D. and Schembri, J. (eds.) The development process in small island states, pp.247-266. London and New York: Routledge, 1993;

SMYTH, W.H. Plan of the harbours and fortifications of Valletta in the island of Malta. London: Hydrogeographical Office, 1821. SOULI, S. Griechische Mythologie. Athens: Michalis Toubis S.A, 1995.

SCHERMERHORN, E.W. Malta of the Knights. Kingswood: W.N. Heinemann Ltd, 1929.

SPRATT, T.A.B. On the Geology of the Maltese Islands. Second edition. Malta: Mail Office, 1854.

SCHMIDT, M. W.; NOACK, A. G. Black carbon in soils and sediments: Analysis, distribution, implications, and current challenges. Global Biogeochemical Cycles, 2000, Vol. 14(3), pp. 777-793.

STAFFORD, K.; MYLROIE, J.; TABOROSI, D.; JENSON, J.; MYLROIE, J. Karst development on Tinian, Commonwealth of the Northern Mariana Islands: controls on dissolution in relation to the carbonate island karst model. Journal of Cave and Karst Studies, 2005, Vol. 67 (1), pp.14-27.

SCHMIDT, M.W.; SKJEMSTAD, J.O.; CZIMCZIK, C. I.; GLASER, B.; PRENTICE, K. M.; GELINAS, Y., KUHLBUSCH, T.A.J. Comparative analysis of black carbon in soils. Global Biogeochemical Cycles, 2001, Vol. 15(1), pp. 163-167.

STANLEY, J.D., HAIT, A.K. Deltas, radiocarbon dating, and measurements of sediment storage and subsidence, Geology, 2000, Vol. 28(4), pp. 295-298.

SCHOCH, W. H.; PAWLIK, B.; SCHWEINGRUBER, F. H. Botanische Makroreste. Berne, Stuttgart: Paul Haupt Publishers, 1988.

STEVENSON, J.; HABERLE, S. Macro Charcoal Analysis: A modified technique used by the Department of Archaeology and Natural History. Palaeoworks Technical Papers 5. URL: Australian National University, 2005.

SCHÖNFELDER, P.; SCHÖNFELDER, I. Was blüht am Mittelmeer? Stuttgart: Frankh-Kosmos, 1990. SCHWAB, M. J.; NEUMANN, F.; LITT, T.; NEGENDANK, J. F.W.; MORDECHAI, S. Holocene palaeoecology of the Golan Heights (Near East): investigation of lacustrine sediments from Birkit Ram crater lake. Quaternary Science Review, 2004, Vol. 23, pp. 1723-1731.

STOCKS, B.J.; KAUFFMANN, J.B. Biomass Consumption and Behaviour of Wildland Fires in Boreal, Temperate, and Tropical Ecosystems: Parameters Necessary to Interpret Historic Fire Regimes and Future Fire Scenarios, in Clark, J.S.; Cachier, H. and Goldammer, J.G. (eds.) Sediment records of biomass burning and global change, pp. 169-188. Berlin Heidelberg: Springer, 1997.

SEILER, W.; CRUTZEN, P.J. Estimates of gross and net fluxes of carbon between the biosphere and the atmosphere from biomass burning. Climatic Change, 1980, Vol. 2, pp. 207-247.

STODDART, S. Mortuary Customs in Prehistoric Malta. In Mifsud, A., Savona Ventura, C. (eds.) Facets of Maltese Prehistory 1999. Malta: Prehistoric Society of Malta, 1999, pp.183-190.

SHARP, M.J.; DOWDESWELL, J.A. Mineral magnetic properties of valley-fill sediments: Implications for provenance and weathering history, in Preece, R.C. and Bridgland, D.R. (eds.), Late Quaternary Environmental Change in North-west Europe. Excavations at Holywell Coombe, South-east England, pp. 86-105. London: Chapman and Hall, 1998.

STOEGER, H. Albert Mayr (1868-1924). Archaeological Review, 2000, Issue 4, pp. 3-9.

Malta

STORCH, G. Holozäne Kleinsäugerfunde aus der Ghar Dalam Höhle, Malta (Mammalia: Insectivora, Chiroptera, Rodentia). Senckenbergiana Biologica, 1970, Vol. 51, pp.135-145. Quartäre Fledermaus-Faunen von der Insel Malta. Senckenbergiana lethaea, 1974, Vol. 55, pp. 407-434.

SINGH, I.B.; JAISWAL, M.; SINGHVI, A.K.; SINGH, B.K. Rapid subsidence of western Ganga plain during late Pleistocene: Evidence from optical dating of subsurface samples. Current Science, 2003, Vol. 84 (3), pp. 451-454.

SUGITA, S.; MACDONALD, G.M.; LARSEN, C.P.S. Reconstruction of Fire Disturbance and Forest Succession from Fossil Pollen in Lake Sediments: Potential and

SKEATES, R.; WHITEHOUSE, R. (eds.). Radiocarbon Dating and Italian Prehistory. London: The British 153

Limitations, in Clark, J.S., Cachier, H. and Goldammer, J.G. (eds.) Sediment records of biomass burning and global change, pp. 387-412. Berlin-Heidelberg: Springer, 1997.

Malta: An Archaeological Guide. Revised edition. Malta: Progress Press, 2000. Malta, Prehistory and Temples. Malta: Midsea Books Ltd, 2002.

SULTANA, J.; FALZON, V. (eds.). Wildlife of the Maltese Islands. Floriana: Environment Protection Department, 1996.

TWIDALE, C.R.; VIDAL ROMANI, J.R.; CAMPBELL, E.M.; CENTENO, J.D. Sheet fractures: response to erosional offloading or to tectonic stress? Zeitschrift für Geomorphologie, Neue Folge, 1996, Supplement Vol. 106, pp. 1-24.

SUN, X.J.; LI, X. A Pollen Record of the Last 37ka in Deep Sea Core 17940 from the Northern Slope of South China. Marine Geology, 1999, Vol. 156 (1-4), p. 227.

TYSON, R.V. Sedimentary Organic Matter. London: Chapman and Hall, 1995.

SWITZER, A.D. Depositional characteristics of recent and late Holocene overwash sandsheets in coastal embayments from southeast Australia. PhD dissertation, University of Wollongong, 2005.

UMBANHOWAR, C. E. Jr.; MCGRATH, M. J. Experimental production and analysis of microscopic charcoal from wood, leaves and grasses. The Holocene, 1998, Vol. 8 (3), pp.341-346.

TAYLOR, G.H.; TEICHMÜLLER, M.; DAVIS, A.; DIESSEL, C.F.K.; LITTKE, R.; ROBERT, P. Organic Petrology. Stuttgart: Gebrüder Bornträger, 1985.

VANCE, J.G. Description of an ancient Temple near Qrendi, Malta. Archaeologia, 1842, Vol. XXIX, pp. 227-40.

TESTA, M. Rough sea stops 25 ships outside ports. The Times of Malta, 2/2/2006.

VASSALLO, C. Guida al Museo ovvero i monumenti di antichità Maltesi conservati nel Museo, descritti dal Dr. Cesare Vassallo. Valletta, 1971.

THAKE, M.A. Land snails from the Mellieha Quaternary Deposit. In Potamon (Malta), 1985, Vol. 14, p.93.

VEGETIUS RENATUS, F. De Re Militari. Antwerp: Christoph. Platinus, 1585.

THOMPSON, R.; OLDFIELD, F. Environmental Magnetism.London: Allen and Unwin, 1986.

VELLA, H.C.R. The Earliest Description of Malta (Lyons 1536) by Jean Quintin d’Autun. Malta: Interprint Ltd, 1980. The Island of Gozo in classical texts. Occasional Papers on Islands and Small States, Vol. 13. Malta: Islands and Small States Institute of the Foundation for International Studies, 1995.

THORNE, C. E. Effects of Vegetation on Riverbank Erosion and Stability. In Thornes, J.B. (ed.), Vegetation and Erosion, pp. 123-144. Chichester: John Wiley and Sons Ltd, 1990. THUCYDIDES. Thucydidis historiae, Vol. 3. Rome: Biblioteca di studi superiori, 1952.

VELLA, N.C. The lie of the land: Ptolemy's temple of Hercules in Malta. Ancient Near Eastern Studies, 2002, Vol. 38, pp. 53-82.

TINNER, W.; HUBSCHMID, P.; WEHRLI, M.; AMMAN, B.; CONDERA, M. Long-term forest fire ecology and dynamics in southern Switzerland. Journal of Ecology, 1999, Vol. 87, pp. 273-289.

VELLA, S. J. Soil Survey and Soil Mapping in the Maltese Islands: the 2003 Position. European Soil Bureau, Research Report No. 9, 2003, pp. 235-244.

TINNER, W.; HU, F. S. Size parameters, size-class distribution and area-number relationship of microscopic charcoal: relevance for fire reconstruction. The Holocene, 2003, Vol 13 (4), pp. 499-505.

VENICE SYSTEM. Final resolution of the symposium on the classification of brackish waters. Archo Oceanogr. Limnol., 1959, Vol. 11 (suppl), pp. 243–248.

TORPIANO, A. Research on Limestone Masonry Structures in Malta. Proc. Int. Seminar, Evoluzione nella Sperimentazioneper le Costruzione, CIAS, Malta, 1996.

VIDAL GONZALEZ, P. La Isla de Malta en Época Fenicia y Púnica. BAR International Series, Vol. 653. Oxford: BAR Publishing, 1996.

TRECHMANN, C.T. Quaternary Conditions in Malta. The Geological Magazine, 1938, Vol. 75, pp. 1-26.

VOSSMERBÄUMER, H. Malta. Ein Beitrag zur Geologie und Geomorphologie des Zentralmediterranen Raumes. Würzburger Geographische Arbeiten, 1972, Vol. 38, pp. 1-212.

TRUMP, D. H. Skorba. London: The Society of Antiquaries, 1966. Malta: An Archaeological Guide. Malta: Progress Press, 1990. Radiocarbon dates from Malta. The Accordia Research Papers, 1995-96, Vol. 6, pp. 173-177.

WAINWRIGHT, J. History and evolution of Mediterranean desertification. Advances in Environmental Monitoring and Modelling, 2004, Vol. 1 (4), pp. 1–87. 154

WARBURTON, J.; HOLDEN, J.; MILLS, A.J. Hydrological controls of surficial mass movements in peat. Earth Science Reviews, 2004, Vol. 67, pp. 139-156.

Avian Representations from prehistoric and medieval sites on the Maltese Islands, in Malta Archaeological Review, 2001, Issue 5, pp.23-39.

WELTER-SCHULTES, F.W. Human-dispersed land snails in Crete, with special reference to Albinaria (Gastropoda: Clausiliidae). Biologia Gallo-hellenica, 1998, Vol. 24 (2), pp. 83-106. Athens.

ZIELHOFER, C.; FAUST, D.; BAENA ESCUDERO, R.; DIAZ DEL OLMO, F.; KADEREIT, A.; MOLDENHAUER, K.-M.; PORRAS, A. Centennialscale late-Pleistocene to mid-Holocene synthetic profile of the Medjerda Valley, northern Tunisia. The Holocene, 2001, 14(6), pp. 851-861.

WENT, D. J. Pre-vegetation alluvial fan facies and processes: an example from the Cambro-Ordovician Rozel Conglomerate Formation, Jersey, Channel Islands. Sedimentology, 2005, Vol. 52 (4), pp. 693-713. WETTINGER, G. Lost Villages and Hamlets of Malta, in Anthony Luttrell (ed.). Medieval Malta: Studies on Malta before the Knights. London: The British School at Rome, 1975, pp. 181-216. Agriculture in the late Middle Ages in Malta, in Proceedings of History Week, 1981, pp. 1-48, Malta: The Historical Society. The Arabs in Malta, in Malta, Study of its Heritage and History, pp.87-104. Malta: Mid-Med Bank Ltd, 1986. Maltese Placenames. Civilization, 1987, Vol. 35, pp. 1489. Aspects of daily life in late-Medieval Malta and Gozo, in Fiorini, S.; Mallia-Milanes, V. (eds.), Malta: A Case Study in International Cross-Currents, pp. 81-90. Malta: University of Malta, 1991 WHITLOCK, C.; MILLSPAUGH, S.H. Testing the assumptions of fire history studies: An examination of modern charcoal accumulation in Yellowstone National Park, USA. Holocene, 1996, Vol. 6, pp. 7-15. WHITLOCK, C.; BRADURY, J.P.; MILLSPAUGH, S.H. Controls on Charcoal Distribution in Lake Sediments: Case Studies from Yellowstone National Park and Northwestern Minnesota, in Clark, J.S., Cachier, H. and Goldammer, J.G. (eds.) Sediment records of biomass burning and global change, pp. 367-386. BerlinHeidelberg: Springer, 1997. WORKING REPORTS OF GOVERNMENT DEPARTMENTS. Report on the working of the Museum Department for the year 1993. Malta: Department of Information, 1993. ZAMMIT, T. Prehistoric Malta. The Tarxien Temples. Oxford: Oxford University Press, 1930. ZAMMIT MAEMPEL, G. The Pleistocene Deposit at Mriehel – Report on the Findings in the South Trench. MAR, 1965, Appendix II, 1965, pp. 8-10. An outline of Maltese Geology. Progress Press: Malta, 1977. A Maltese Pleistocene sequence capped by volcanic tufa. Atti della Società Toscana di Scienze Naturali Residente in Pisa. 1982, Serie A 88, pp. 243-260. Ghar Dalam cave and deposits. Malta: PEG, 1989.

155