[go: up one dir, main page]

Academia.eduAcademia.edu
Metadata of the chapter that will be visualized online Chapter Title Harbors and ports, ancient Copyright Year 2015 Copyright Holder Springer Science+Business Media Dordrecht Corresponding Author Family Name Marriner Particle Given Name Nick Suffix Author Division/Department CNRS, Laboratoire ChronoEnvironnement UMR 6249 Organization/University Université de Franche-Comté Street UFR ST, 16 route de Gray Postcode 25030 City Besançon Country France Email nick.marriner@univ-fcomte.fr Email marriner@cerege.fr Family Name Morhange Particle Given Name Christophe Suffix Author Organization/University Aix-Marseille Université, IUF, CEREGE UMR 6635 Street Europôle de l’Arbois, BP 80 Postcode 13545 City Aix-en-Provence cedex 04 Country France Family Name Flaux Particle Given Name Clément Suffix Author Organization/University Aix-Marseille Université, IUF, CEREGE UMR 6635 Street Europôle de l’Arbois, BP 80 Postcode 13545 City Aix-en-Provence cedex 04 Country France Family Name Carayon Particle Given Name Nicolas Suffix Organization/University Archéologie des Sociétés Méditerranéennes, UMR 5140 Street 390 avenue de Pérols Postcode 34970 City Lattes Country France Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:07 Page Number: 1 Title Name: EOG H 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 Scandinavia (Ilves, 2009), the Mediterranean (Marriner and Morhange, 2007), and Africa (Chittick, 1979). HARBORS AND PORTS, ANCIENT 1 2 34 35 2 Nick Marriner , Christophe Morhange , Clément Flaux and Nicolas Carayon3 1 CNRS, Laboratoire Chrono-Environnement UMR 6249, Université de Franche-Comté, Besançon, France 2 Aix-Marseille Université, IUF, CEREGE UMR 6635, Aix-en-Provence cedex 04, France 3 Archéologie des Sociétés Méditerranéennes, UMR 5140, Lattes, France Synonyms Haven; Port; Roadstead Definition Coastal areas have been used as natural roadsteads at least since prehistoric times. In the Oxford English dictionary, a harbor is “a place on the coast where ships may moor in shelter, especially one protected from rough water by piers, jetties, and other artificial structures.” This safe refuge can be either natural or artificial. As a result, the term “harbor” can often be ambiguous when it refers to a premodern context because it incorporates a plethora of landing site types, including offshore anchorages, in addition to different mooring facilities and technologies (Raban, 2009). Conceptions of ancient Mediterranean harbors have frequently been skewed by all-season harbor facilities such as Alexandria, Piraeus, and Valletta with their favorable geomorphological endowments. The archaeological record is, however, more complex. Port is derived from the Latin portus meaning “opening, passage, asylum, refuge.” Drawing on multidisciplinary archaeological and geoscience tools, there has been a renewed interest in ancient harbors during the past 30 years, including the Indian Ocean (Rao, 1988), the Atlantic, Introduction Until recently, coastal sediments uncovered during Mediterranean excavations received very little attention from archaeologists, even though, traditionally, the received wisdom of Mare Nostrum’s history has placed emphasis on the influence and coevolution of physical geography in fashioning its coastal societies (Braudel, 2002; Stewart and Morhange, 2009; Martini and Chesworth, 2010; Abulafia, 2011). Before 1990, the relationships between Mediterranean populations and their coastal environments had been studied within a cultural-historical paradigm, where anthropological and naturalist standpoints were largely considered in isolation (Horden and Purcell, 2000). During the past 20 years, Mediterranean archaeology has changed significantly, underpinned by the emergence of a new culture-nature duality that has drawn on the North European examples of wetland and waterfront archaeology (Milne and Hobley, 1981; Coles and Lawson, 1987; Purdy, 1988; Coles and Coles, 1989; Mason, 1993; Van de Noort and O’Sullivan, 2006; Menotti and O’Sullivan, 2012). This built on the excavation of Alpine lake settlements in Switzerland and elsewhere from the 1850s onwards (Keller, 1866). Because of the challenges of waterfront contexts, the archaeological community is today increasingly aware of the importance of the environment in understanding the socioeconomic and wider natural frameworks in which ancient societies lived, and multidisciplinary research and dialogue have become a central pillar of most large-scale excavations (Walsh, 2004; Butzer, 2005; Butzer, 2008; Walsh, 2008). It is against this backdrop that ancient harbor contexts have emerged as particularly novel archives, shedding new light on how humans have locally interacted with and modified coastal zones since the Neolithic (Marriner A.S. Gilbert (ed.), Encyclopedia of Geoarchaeology, DOI 10.1007/978-1-4020-4409-0, © Springer Science+Business Media Dordrecht 2015 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:07 Page Number: 2 2 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 Title Name: EOG HARBORS AND PORTS, ANCIENT and Morhange, 2007). Their importance in understanding ancient maritime landscapes and societies (e.g., Gambin, 2004; Gambin, 2005; Tartaron, 2013) makes them one of the most discussed archaeological contexts in coastal areas (Figure 1). Around 6,000 years ago, at the end of the Holocene marine transgression, societies started to settle along “present” coastlines (Van Andel, 1989). Older sites were buried and/or eroded during this transgression (Bailey and Flemming, 2008). During the past 4,000 years, harbor technology has evolved to exploit a wide range of environmental contexts, from natural bays and estuaries through to the completely artificial basins of the Roman and Byzantine periods. Although some of these ancient port complexes continue to be thriving transport centers, now, many millennia after their initial foundation, the vast majority have been completely abandoned, and their precise whereabouts, despite rich textual and epigraphic evidence, remain unknown. Although not the sole agent of cultural change, these environmental modifications indicate in part that long-term human subsistence has favored access to the open sea. Key to this line of thinking is the idea that societies have adopted adaptive strategies in response to the rapidly changing face of the coastal environment, and in many instances, harbor sites closely mirror modifications in the shoreline (e.g., Brückner et al., 2004). Nonetheless, it is important to emphasize that regional environmental change, although strong, must not be seen as the principal agent of cultural shifts and that site-specific explanations remain fundamental (Butzer, 1982). During the 1960s, urban regeneration led to large-scale urban excavations in many coastal cities of the Mediterranean. It was at this time that the ancient harbor of Marseille (France) was rediscovered. Nonetheless, it was not until the early 1990s that two large-scale coastal excavations were undertaken at opposite ends of the Mediterranean in Marseille (Hesnard, 1994; Hesnard, 1995) and Caesarea Maritima in Israel (Raban and Holum, 1996). Both projects placed emphasis on the harbor archaeology and their articulation within the wider landscape. The first, at Caesarea Maritima, investigated a completely artificial Roman harbor complex on the Levantine coast, active between the first and second centuries AD (Reinhardt et al., 1994; Reinhardt and Raban, 1999; Raban, 2009). At Marseille, meanwhile, researchers set about reconstructing the archaeology and environmental history of the city’s ancient harbor since the seventh century BC, founded in a naturally protected limestone embayment by Greek colonists from Ionia (Figure 2). In contrast to deltaic areas, the smaller analytical scale of harbor basins meant that coastal changes could be studied not only with greater facility but also more finitely. The research at Marseille (Morhange et al., 2003) reconstructed a rapid shift in shoreline positions from the Bronze Age onwards and demonstrated the type of spatial resolution that can be obtained when large excavation areas are available for geoarchaeological study. These studies were unique in that, for the first time in a Mediterranean coastal context, both sought to embrace a multidisciplinary methodology. Investigative fields included not only archaeology but also geomorphology, geography, sedimentology, history, and biology (Raban and Holum, 1996; Hesnard, 2004). The waterlogged conditions were particularly conducive to environmentally contextualized analyses, and both studies demonstrated how coastal archaeology could benefit from being placed within a broader multidisciplinary framework. Since these projects, there has been a great proliferation of studies looking into coastal and ancient harbor geoarchaeology (see Marriner and Morhange, 2007 for multiple references; Figure 1), building on pioneering archaeological work in the first half of the twentieth century (e.g., Negris, 1904a; Negris, 1904b; Paris, 1915; Jondet, 1916; Paris, 1916; Lehmann-Hartleben, 1923; Poidebard, 1939; Halliday Saville, 1941; Poidebard and Lauffray, 1951). Ancient harbor basins are particularly interesting because (1) they served as important economic centers and nodal points for maritime navigation (Casson, 1994; Arnaud, 2005); (2) there is generally excellent preservation of the material culture (Rickman, 1988; Boetto, 2012) due to the anoxic conditions induced by the water table; and (3) there is an abundance of source material for paleoenvironmental reconstruction (Marriner, 2009). Seaports are particularly interesting, as they allow us to understand how people “engaged with” the local environmental processes in coastal areas. Here, we will explore the specific interest of harbor sediments in reconstructing ancient coastal landscapes and their evolution through time. In particular, we will discuss the stratigraphic evidence for these changes and set them within the wider context of coastal changes driven by various natural and anthropogenic forcing agents. We will also address present challenges and gaps in knowledge. 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 Harbor origins The ease of transport via fluvial and maritime routes was important in the development of civilizations. At least three areas – the Indus, China, and Egypt – played an important role in the development of harbors and their infrastructure. 163 Egypt It has been suggested that the Egyptians were one of the earliest Mediterranean civilizations to engage in fluvial and maritime transportation. Evidence for the use of boats in ancient Egypt derives from deepwater fish bones found at prehistoric hunter/gatherer campsites (Shaw et al., 1993). The earliest boats were probably rafts made of papyrus reeds, which enabled these societies to navigate between camps. It is speculated that wooden boats were adopted during Neolithic times, around the same time as the introduction of agriculture and animal husbandry. The rise of chiefdoms during the Egyptian Predynastic period (3700–3050 BC) was accompanied by the widespread adoption of boats as attested by art and pottery 169 164 165 166 167 168 170 171 172 173 174 175 176 177 178 179 180 181 182 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:07 Page Number: 3 Title Name: EOG HARBORS AND PORTS, ANCIENT 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 depictions (Fabre, 2004–2005). North of the First Cataract in Egypt, ships could travel almost anywhere along the Nile. On the delta, the then seven branches served as navigable waterways into the Eastern Mediterranean (Tousson, 1922; Stanley, 2007; Khalil, 2010). The Eastern Mediterranean was also a natural communications link for the major cultural centers of the Levant, Cyprus, Crete, Greece, and North Africa. In light of this, it is unsurprising that the works along the fluvial banks and coastlines of the Red Sea and Mediterranean were many and varied. During the third millennium BC, canals were excavated from the Nile to the valley temples of the Giza pyramids so that building materials could be transported (Fabre, 2004–2005; Butzer et al., 2013). Quays were also commonly established along the Nile, for instance, at fourteenth century BC Amarna, boats have been depicted parallel to shoreside quays equipped with bollards (Blackman, 1982a; Blackman, 1982b). An artificial quay dating to the second millennium BC is attested at Karnak, on the Nile (Lauffray et al., 1975; Fabre, 2004–2005). High sediment supply and rapid changes in fluvial systems mean that few conspicuous remains of these early riverine harbors are still visible, particularly on the delta (Blue and Khalil, 2010). In Mesopotamia, a similar evolution is attested (Heyvaert and Baeteman, 2008). Navigation in the Red Sea during pharaonic times is a theme that has attracted renewed interest during the past 30 years, underpinned notably by the discovery of a number of exceptional coastal sites, shedding new light on the extent and chronology of human impacts in maritime areas. Extending for over 2,000 km from the Mediterranean Sea to the Arabian Sea, the Red Sea was a major communications link. Egyptian seafarers traveled along its shorelines during the Predynastic period and were probably the first to contact the peoples living on the Sudanese coast and around the Horn of Africa. Since the discovery of remains at Mersa/Wadi Gawasis in 1976, new findings have been made more recently at Ayn Soukhna, El-Markha, and Wadi al-Jarf (Tallet, 2009). In the absence of harbor excavations, much of the data available remain preliminary. At Mersa/Wadi Gawasis, archaeological data have documented evidence for some of the world’s earliest long-distance seafaring, including bundled ropes, ships, and remnants of storage boxes used for transport of goods. The site was used extensively during the Middle Kingdom (around 4,000–3,775 years ago), when seafaring ships departed from the harbor for trade routes along the African Red Sea coast (Bard and Fattovich, 2010; Hein et al., 2011). The Indus Valley On the Indian subcontinent, archaeological explorations during the past century have brought to light a large number of structures related to ancient harbor works and maritime activities (Rao, 1988). The Indus valley in particular has been a key focus of research, where high sediment supply in a context of rapidly changing deltaic 3 environments is responsible for the landlocking of many ancient port sites (Gaur and Vora, 1999). The oldest reference to a harbor in India derives from a mid-third millennium Mesopotamian text mentioning boats from Meluhha that were anchored in Agade harbor (Kramer, 1964). Nonetheless, despite rich textual evidence, the exact location of many of these ancient harbor sites is equivocal. Most would have exploited riverbanks that served as natural harbors. Many of the best-studied examples derive from the region of Gujarat, which attests to significant paleo-shoreline changes during the past 4,500 years (Gaur and Vora, 1999). Archaeological sites of Harappan age (3000–1500 BC), including Lothal, Padri, and Bet Dwarka, have yielded particularly interesting archaeological records consistent with maritime activity (Gaur and Vora, 1999). Lothal, on the paleo-banks of the river Sabarmati, is one of the best-studied examples of a Harappan harbor city. The site presently lies 35 km from the coast at the head of the macrotidal Gulf of Cambay and is believed to have been an important trade center during the Harappan period (Rao, 1991). A number of Egyptian and Mesopotamian imports have been recovered from the site. Excavations have brought to light a brick basin of trapezoidal shape that measures 214  36 m and is 3.3 m deep. It has tentatively been labeled as the world’s first dockyard (Rao, 1979), although these interpretations are not without contention (e.g., Gaur, 2000), and the basin presents striking similarities with water storage basins used throughout the region. Based on present knowledge, it is difficult to confirm that Lothal’s basin was used as a harbor. Elsewhere in the Indus valley, Chalcolithic/Harappan landing platforms attributed to harbor works have been identified at Kuntasi and Inamgaon. Paleoenvironmental changes are seen as important causes of harbor abandonment. China Between 7000 and 5000 BC, agricultural villages and towns began to emerge and grow along the Yellow and Yangtze River basins and coasts. Research has focused on this transitional period because it corresponds to the onset of deltaic sedimentation and the emergence of agriculture and early complex societies (Zong et al., 2007; Chen et al., 2008). Ancient Chinese history is marked by three successive dynasties that became the roots of Chinese culture: the Xia Dynasty (2200–1766 BC), the Shang Dynasty (1766–1122 BC), and the Zhou Dynasty (1122–256 BC). Despite the importance and continuity of Chinese civilization, understanding of its harbors is relatively limited in western academic circles due to obvious language barriers. Nonetheless, the recent rediscovery of Hepu harbor of the Western Han Dynasty (206 BC to 25 AD) is particularly promising in shedding new light on this question. Now located within Beihai City in south China’s Guangxi Zhuang Region, recent archaeological work suggests that Hepu harbor – probably the oldest seaport in China – served as a very important “marine silk 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:08 Page Number: 4 4 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 Title Name: EOG HARBORS AND PORTS, ANCIENT road.” This navigation link allowed western goods to be transported into the vast continental interior of Asia. Early Mediterranean harbors Our understanding of early harbors is poor. In the Mediterranean, the first artificial structures appear to date to the Middle/Late Bronze Age. For example, submerged boulder piles are attested at Yavne-Yam, a Middle Bronze Age site on the coast of Israel; these suggest premeditated human enterprise to improve the quality of the natural anchorage (Ezra Marcus, personal communication). Recent geoarchaeological work in Sidon (Lebanon) has tentatively dated the presence of a semi-protected cove beginning around 4410  40 BP (2750–2480 cal BC; Marriner et al., 2006b; Marriner, 2009). This sedimentological unit has been interpreted as a Middle Bronze Age to Late Bronze Age proto-harbor, with possible reinforcement of the shielding sandstone ridge improving the quality of the natural anchorage. It is suggested that small boats were beached, with larger vessels being anchored in the outer harbor of Zire (Frost, 1973; Carayon, 2008; Figure 3). At Kommos, in southern Crete, a large building with six galleries (Puglisi, 2001) has been interpreted as a hangar for the dry-docking of Minoan ships during the winter months. This building, dated to the fifteenth century BC, is an illustration of Minoan harbor construction even though, in this instance, it had no direct impact upon the quality of the anchorage haven. After this period, the maritime harbors of the ancient Mediterranean evolved in four broad technological leaps. Bronze Age to early Iron Age ashlar header technology A double ashlar wall infilled with stones is a harbor construction method common to the Phoenicians; it is known as the pier-and-rubble technique (Raban, 1985). This system has been noted in an eleventh century BC layer at Sarepta, Lebanon (Markoe, 2000). Van Beek and Van Beek (1981) have suggested that this technique is Levantine in origin and that it spread from the Late Bronze Age Levant to the western Punic colonies, Greece, and Roman North Africa, where it can be found as late as the sixth century AD. The use of ashlar techniques is well attested in the Persian period harbor of Akko (Israel), the Hellenistic harbor at Amathus in Cyprus (Empereur and Verlinden, 1987), and the Roman quay at Sarepta, Lebanon (Pritchard, 1978), Dor, and Athlit (Israel). Iron Age Athlit is one of the best-studied Phoenician harbors (Haggi, 2006; Haggi and Artzy, 2007). The northern harbor’s mole extends about 100 m into the sea. It is about 10 m wide and constitutes two parallel ashlar headers that are 2–3 m in width. A fill of rubble and stones was placed between the ashlar walls. This form of construction improved the stability of the mole against high-energy waves. The mole was placed on a foundation of ballast pebbles of various sizes. Underwater excavations have revealed that the layer of pebbles extends more than 5 m beyond the outer side of each wall, a total width of over 20 m. Radiometric dating of wood fragments constrains this Phoenician structure to the ninth century BC (Haggi, 2006), although paradoxically there is very little pottery dating from this period (Michal Artzy, personal communication). A similar example is also known from the Syrian coast at Tabbat el-Hammam, where the archaeological evidence supports a ninth/eighth century BC age (Braidwood, 1940). Depending on the time and culture, different variations are noted in the use of headers. From the fifth century BC, metal links were used to reinforce blocks (e.g., Sidon and Beirut). At Amathus (Cyprus) during Hellenistic times, the header masonry was built upon a ballast base of disorganized blocks. Cothons Archaeologists refer to the sites of Carthage (Tunisia), Mahdia (Tunisia), Phalasarna (Crete), Jezirat Fara’un (Egypt), and Lechaion (Greece) as “cothon” harbors. The Greek term was applied to the harbor at Carthage by Strabo and Appian, the original meaning of “drinking cup” which is metaphorically appropriate to the protected harbor basin. Carthage is the only site that has been referred to as a “cothon” in ancient texts, although a Punic etymology has not yet been supported, meaning it is difficult to propose that the concept was Carthaginian in origin or that all harbors built into the shoreline in the same manner were felt to be variations on a “cothon” (John Oleson, personal communication). Nowadays, specialists agree that the term can be associated with an artificially dug harbor basin linked to the sea via a man-made channel (Carayon, 2005). The design solves some of the problems involved in building a harbor along a shallow, featureless coastline, or on the bank of a river, and a number of cultures appear to have adopted this solution, from the Bronze Age onwards. Some authors have suggested that Trajan’s basin at Portus also qualifies as a cothon, in addition to some of the proposed Etruscan harbor basins associated with river mouths (John Oleson, personal communication). It would appear that the carving of a cothon is a simple but energy-consuming technique used to create a particularly well-sheltered basin. This type of infrastructure poses three problems: (1) rapid silting up in a confined environment; (2) the carving of a basin in rocky outcrops or clastic coastlines, which is energy consuming; and (3) maintaining a functional channel outlet to the sea in a clastic coast context. Despite these shortcomings, the cothon persisted for many centuries (Carayon, 2008). A Latin author, writing in the fifth century AD, noted that this type of harbor was common at this time: “ut portus scilicet faciunt” (Deutero-Servius, Aeneidos, I, 421). 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:08 Page Number: 5 Title Name: EOG HARBORS AND PORTS, ANCIENT 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 Hydraulic concrete Pre-Roman ashlar block methods continued to be used throughout the Roman era. Nonetheless, another technique was introduced during the second century BC (Gazda, 2001) that completely revolutionized harbor design and construction – the use of hydraulic concrete. This technological breakthrough meant that natural roadsteads were no longer a prerequisite to harbor loci, and completely artificial ports, enveloped by imposing concrete moles, could be located on open coasts (Hohlfelder, 1997). The material could be cast and set underwater. Roman architects and engineers were free to create structures in the sea or along high-energy shorelines (Brandon et al., 2005; Brandon et al., 2010). Pozzolana facilitated the construction of offshore basins such as Claudius’s harbor at Portus of Rome (Testaguzza, 1970). The Roman author Vitruvius (first century BC) provided an inventory of harbor construction techniques (Vitruvius, De Architectura, V, 12). Romano-Byzantine harbor dredging Vitruvius gave a few brief accounts of dredging, although direct archaeological evidence has, until now, remained elusive. The ancient harbors of Marseille and Naples have both undergone widespread excavations (Figure 4; Hesnard, 1995; Giampaola et al., 2004), and extensive multidisciplinary datasets now exist for the two sites. At Tyre and Sidon, geoarchaeological research has led to the extraction of 40 cores that have facilitated a chronostratigraphic reconstruction of basin silting (Marriner et al., 2005; Marriner and Morhange, 2006a; Morhange and Marriner, 2010a). Why were ancient harbors dredged? On decadal timescales, continued silting induced a shortening of the water column. De-silting infrastructure (Blackman, 1982a; Blackman, 1982b), such as vaulted moles, partially attenuated the problem, but in the long term, these appear to have been relatively ineffective. In light of this, repeated dredging was the only means of maintaining a practicable draft depth and ensuring longterm harbor viability. At Marseille, although dredging phases are recorded from the third century BC onwards, the most extensive enterprises were undertaken during the first century AD, at which time huge volumes of sediment were extracted. At the excavations of Naples, absence of pre-fourth century BC layers has been linked to extensive dredging between the fourth and second centuries BC (Carsana et al., 2009). Unprecedented traces 165–180 cm wide and 30–50 cm deep attest to powerful dredging technology that scoured into the volcanic substratum, completely reshaping the harbor bottom. Notwithstanding the scouring of harbor bottoms, this newly created space was rapidly infilled and necessitated regular intervention. Repeated dredging phases are attested up until the late Roman period, after which time the basin margins were completely silted up. At Marseille, three dredging boats have been unearthed (Pomey, 1995). The vessels were abandoned at the bottom of the harbor during 5 the first and second centuries AD. They are characterized by an open central well that is inferred to have accommodated the dredging arm. It was not until the Industrial Revolution in England that cement and iron structures were developed on a large scale (Palley, 2010). In 1756, Smeaton made the first modern concrete (hydraulic cement) by adding pebbles as a coarse aggregate and mixing powdered brick into the cement. In 1824, Aspdin invented Portland cement by burning ground limestone and clay together. The Frenchman Monier invented reinforced concrete in 1849 using imbedded steel. It can withstand heavy loads because of its tensile and compressional strengths. Reinforced concrete was widely used in railway ties, pipes, floors, arches, bridges, and ports. 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 Geoarchaeology of harbor basins: tools and methods Over the past 2 decades, ancient harbors have attracted interest from both the archaeological and earth science communities. In tandem with the development of rescue archaeology, particularly in urban contexts, the study of sedimentary archives has grown into a flourishing branch of archaeological inquiry (Milne, 1985; Leveau et al., 1999; Milne, 2003; Walsh, 2004; Leveau, 2005). The growing corpus of sites and data demonstrates that ancient harbors constitute rich archives of both the cultural and environmental pasts. Ancient harbor sediments are particularly rich in research objects (archaeological remains, bioindicators, macrorests, artifacts, etc.), and they yield insights into the history of human occupation at a given site, coastal changes, and the natural processes and hazards that have impacted these waterfront areas (Reinhardt et al., 2006; Bottari and Carveni, 2009; Morhange and Marriner, 2010b; Bony et al., 2012). Ancient harbors are both natural and constructed landscapes and, from a geoarchaeological perspective, comprise three elements of note. 474 The harbor basin In architectural terms, the harbor basin is characterized by its artificial structures, such as quays, moles, and sluice gates (Oleson, 1988; Oleson and Branton, 1992). Since the Bronze Age, there has been a great diversity in harbor infrastructure in coastal areas, reflecting changing technologies and human needs. These include, for instance, the natural pocket beaches serving as proto-harbors (Frost, 1964; Marcus, 2002a; Marcus, 2002b), through the first Phoenician mole attributed to around 900 BC (Haggi and Artzy, 2007), to the grand offshore constructions of the Roman period made possible by the discovery of hydraulic concrete (Oleson et al., 2004). In their study of harbor landscapes, geoarchaeologists are also interested in the sedimentary contents of the basin and relative sea-level changes. 496 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:08 Page Number: 6 6 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 Title Name: EOG HARBORS AND PORTS, ANCIENT Ancient harbor sediments Port basins constitute unique coastal archives. Shifts in the granularity of these deposits indicate the degree of harbor protection, often characterized by a rapid accumulation of heterometric sediments following a sharp fall in water competence brought about by the installation of artificial harbor works. The harbor facies is characterized by three poorly sorted fractions: (1) human waste products, especially at the base of quays and in areas of unloading (harbor depositional contexts are particularly conducive to the preservation of perishable artifacts such as leather and wood); (2) poorly sorted sand; and (3) an important fraction (>90 %) of silt that signifies the sheltered environmental conditions of the harbor. They are also particularly pertinent archives for reconstructing the history of heavy metal pollution at coastal settlements (e.g., Véron et al., 2006). Harbor basins are characterized by rapid accumulation rates. For instance, sedimentation rates of up to 20 mm/year have been recorded in undredged areas of the Graeco-Roman harbor of Alexandria (Goiran, 2001). High-resolution study of the bio- and lithostratigraphical fractions can help shed light on the nature of ancient harbor works, such as at Tyre (Marriner et al., 2008) or Portus (Goiran et al., 2010). Recent research has sought to characterize and date these chronostratigraphic phases using the unique sedimentary signature that each technology brings about (Marriner and Morhange, 2007; Marriner, 2009). In the broadest sense, these are characterized by an evolution from natural roadsteads before the Bronze Age towards completely artificial seaport complexes from the Roman period onwards. Relative sea-level changes, the paleo-water column, and ship circulation Nowadays, most ancient harbors are completely infilled with sediments – e.g., the Roman harbor of Luni at the mouth of the river Magra (Bini et al., 2009) or the Roman harbor of Aquileia (Arnaud-Fassetta et al., 2003). Harbor sediments are particularly conducive to the preservation of biological remains. Within this context, it is possible to identify and date former sea-level positions using biological indicators fixed to quays, that, when compared with the marine bottom, allow the height of the paleowater column to be estimated (Laborel and LaborelDeguen, 1994; Morhange et al., 2013). Such relative sea-level data are critical in understanding the history of sedimentary accretion in addition to estimating the draft depth for ancient ships (Pirazzoli and Thommeret, 1973; Morhange et al., 2001; Boetto, 2012). Archaeological work undertaken upon ancient wrecks suggests that the largest fully loaded ships during antiquity required a draft of less than 3 m (Casson, 1994; Pomey and Rieth, 2005). These two reference levels, the paleo-sea level and sediment bottom, are mobile as a function of crustal movements – e.g., local-scale neotectonics (Stiros et al., 1996; Stiros, 1998; Evelpidou et al., 2011), regional isostasy (Lambeck et al., 2004), sediment budgets (Vött et al., 2007; Devillers, 2008), and human impacts such as dredging (Marriner and Morhange, 2006b). All these factors can potentially impact the available accommodation space for sediment accretion. 568 569 570 571 Sediments versus settlements As outlined above, one of the key problems posed by artificially protected harbors relates to accelerated sediment trapping. In the most acute instances, it could rapidly reduce the draft depths necessary in accommodating large ships (Pomey and Rieth, 2005). From a cultural perspective, therefore, harbors were important “economic landscapes,” and many changes in harbor location can be explained functionally by the need to maintain an interface with the sea in the face of rapid sedimentation. The best example of this coastal dislocation derives from Aegean Anatolia (Brückner et al., 2005). Delta areas in particular serve as excellent geo-archives to understand and analyze the impacts of rapidly evolving settlement phases. It is important to set these geoarchaeological results within a wider spatiotemporal framework using archaeological data from coastal and hinterland valley areas. Changes in sediment supply at the watershed scale are particularly important in understanding base-level changes in deltaic and coastal contexts, as is the case of the Gialias in Cyprus (Devillers, 2008) or the paleo-island of Piraeus (Goiran et al., 2011). Probing the rates of progradation is also key to understanding the timing, origin (climate or human forcings), and rhythm of local and basin-scale erosion. 572 Ancient harbor stratigraphy, terminology and research goals During the past 20 years, multidisciplinary inquiry has allowed a better understanding of where, when, and how ancient Mediterranean harbors evolved. This is set within the wider context of a new “instrumental” or “quantitative revolution” towards the environment. A battery of research tools is available, tools that broadly draw on geomorphology and the sediment archives located within this landscape complex (Marriner and Morhange, 2007). 597 Where? The geography of ancient harbors constitutes a dual investigation that probes both the location and the extension of the basins. Biostratigraphical studies of sediments, married with a GIS investigation of aerial photographs and satellite images, can be used to reconstruct coastal evolution and identify possible anchorage areas (Ghilardi and Desruelles, 2009). Traditionally, urban contexts have been particularly problematic for accurate archaeological studies because the urban fabric can hide many of the most important landscape features. In such instances, chronostratigraphy can be particularly useful in reconstructing coastal changes (Morhange et al., 2003). For example, litho- and biostratigraphical studies of cores drilled into the city center of Tyre attest to a well-sheltered 607 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 598 599 600 601 602 603 604 605 606 608 609 610 611 612 613 614 615 616 617 618 619 620 621 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:08 Page Number: 7 Title Name: EOG HARBORS AND PORTS, ANCIENT 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 port basin between the Hellenistic and Byzantine periods, today buried beneath the modern market by thick sediment tracts. The chronostratigraphy demonstrates that during antiquity, the harbor was approximately twice as large as present (Figure 5). This approach helps not only in reconstructing ancient shorelines and changes through time (e.g., as at Ephesus, Priene, Frejus, Alexandria, or Pelusium on the Nile Delta) but can also aid in relocating ports for which no conspicuous archaeological evidence presently exists, as in the case of Cuma (Stefaniuk and Morhange, 2005) or Byblos (Stefaniuk et al., 2005). Geophysical techniques can also provide a great multiplicity of mapping possibilities, notably in areas where it is difficult to draw clear parallels between the archaeology and certain landscape features (Nishimura, 2001). Because geophysical techniques are nondestructive, they have been widely employed in archaeology and are gaining importance in coastal geoarchaeology (Hesse, 2000) and ancient harbor contexts (Boyce et al., 2009). Very rapid and reliable information can be provided on the location, depth, and nature of buried archaeological features before excavation. At Alexandria, geophysical surveys have allowed Hesse (1998) to propose a new hypothesis for the location of the Heptastadium. Hesse suggests that the causeway linking Pharos to the mainland was directly tied into the city’s ancient road network. In this instance, the findings have since been corroborated by sedimentological data from the tombolo area (Goiran, 2001). Stratigraphic data are therefore critical in providing chronological insights into environmental changes and coastal processes. Such a dual approach has also been successfully employed at Portus, one of the ancient harbors of Rome. Large areas of the seaport and its fringes have been investigated using coastal stratigraphy (Bellotti et al., 2009; Giraudi et al., 2009; Goiran et al., 2010; Di Bella et al., 2011; Mazzini et al., 2011; Salomon et al., 2012), geophysics, and archaeological soundings (Keay et al., 2005; Keay et al., 2009; Keay and Paroli, 2011), yielding fresh insights into the harbor’s coastal infrastructure and functioning. On the Tiber delta, geophysics has also been used to accurately map the progradation of the coastal ridges. Bicket et al. (2009) have demonstrated that the Laurentine ridge, 1 km inland from the modern coastline, constitutes the Roman shoreline of the Tiber delta. When and how? Chronostratigraphy is essential in understanding modifications in harbor technology and the timing of human impacts, such as lead pollution from the Bronze Age onwards (Véron et al., 2006) or ecological stresses demonstrated by changes in faunal assemblages (Leung Tack, 1971–72). The overarching aim is to write a “sedimentary” history of human coastal impacts and technologies, using quantitative geoscience tools and a standardized stratigraphic framework (e.g., sequence stratigraphy). Research in the eastern and western Mediterranean attests to considerable repetition in ancient 7 harbor stratigraphy, both in terms of the facies observed and their temporal envelopes. There are three distinct facies of note: (1) middle-energy beach sands at the base of each unit (e.g., the proto-harbor), (2) low-energy silts and gravels (e.g., the active harbor phase), and (3) coarsening up beach sands or terrestrial sediments which cap the sequences (e.g., post-harbor facies). In the broadest terms, this stratigraphic pattern represents a shift from natural coastal environments to anthropogenically modified contexts, before a semi- or complete abandonment of the harbor basin. There are a number of stratigraphic surfaces that are key to understanding the evolution of ancient harbor basins. The maximum flooding surface (MFS) Ancient harbors form integral components of the highstand parasequence (aggradational to progradational sets). For the Holocene coastal sequence, the maximum flooding surface (MFS) represents the lower boundary of the sediment archive. This surface is broadly dated to around 6000 cal BP and marks the maximum marine incursion (Stanley and Warne, 1994). It is associated with the most landward position of the shoreline. In the eastern Mediterranean, it is contemporaneous with the Chalcolithic period and the Early Bronze Age. Indeed, the MFS along the Levantine coast clearly delineates the geography of early coastal settlements from this period (Raban, 1987). Natural beach facies The MFS is overlain by naturally aggrading beach sands, a classic feature of clastic coastlines. Since around 6000 cal BP, relative sea-level stability has impinged on the creation of new accommodation space, leading to the aggradation of sediment strata. This is particularly pronounced in sediment-rich coastal areas such as deltas and at the margins of fluvial systems. Where this sedimentation continued unchecked, a coarsening upward of sediment facies is observed, consistent with high-energy wave dynamics in proximity to mean sea level. For example, Gaza bears witness to important coastal changes since the Bronze Age. During the mid-Holocene, the coast comprised estuaries at the outlets of major wadi systems. This indented coastal morphology spawned important maritime settlements such as Tell es-Sakan and Tell al-’Ajjul at the outlet of Wadi Ghazzeh, which probably served as a natural harbor. During the same period, the rate of sea-level rise slowed, leading to the formation of the Nile Delta and small, local deltas along the coasts of Sinai and Palestine. From the first millennium BC onwards, the coast was regularized by infilling of the estuaries, and the harbor sites became landlocked. In response, new cities, such as Anthedon, were founded on a Quaternary ridge along the present coastline (Morhange et al., 2005). The harbor foundation surface (HFS) This surface marks important human modification of the sedimentary environment, characterized by the transition 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:09 Page Number: 8 8 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 Title Name: EOG HARBORS AND PORTS, ANCIENT from coarse beach sands to finer-grained harbor sands and silts (Marriner and Morhange, 2007). This surface corresponds to the construction of artificial harbor works and, for archaeologists, is one of the most important surfaces to date the foundation of the harbor. The ancient harbor facies (AHF) The AHF corresponds to the active harbor unit. This artificialization is reflected in the sedimentary record by lower-energy facies consistent with a barring of the anchorage by artificial means. Harbor infrastructure (quays, moles, and jetties) accentuated the sediment sink properties by attenuating the swell and marine currents leading to a sharp fall in water competence. Research has demonstrated that this unit is by no means homogeneous, with harbor infrastructure and the nature of sediment sources playing a key role in shaping facies architecture. Of note is the granulometric paradox of this unit consisting of fine-grained silts juxtaposed with coarse gravels made up of ceramics and other urban waste. In some rare instances, a proto-harbor phase (PHP) precedes the AHF. Before the major changes characteristic of the AHF, biosedimentological studies have elucidated moderate signatures of human presence when societies exploited natural low-energy shorelines requiring little or no human modification. For instance, coastal stratigraphy has demonstrated that the southern cove of Sidon, around Tell Dakerman, remained naturally connected and open to the sea throughout antiquity (Poidebard and Lauffray, 1951; Marriner et al., 2006a; Marriner et al., 2006b). The PHP interface is by no means transparent, particularly in early Chalcolithic and Bronze Age harbors, and the astute use of multiproxy data is required (Figure 6). During the Late Bronze Age and Early Iron Age, improvements in harbor engineering have been recorded by increasingly fine-grained facies. Plastic clays tend to be the rule for Roman and Byzantine harbors, and sedimentation rates 10–20 times greater than naturally prograding coastlines are recorded. The very wellprotected Roman harbors of Alexandria, Marseille, and Frejus (Gébara and Morhange, 2010) all comprise plastic marine muds consisting of 90 % silts and a coarse gravel fraction of human origin. Significant increases in sedimentation rates can also be attributed to human-induced increases in the supply term, for example, anthropogenic changes in the catchments of supplying rivers (deforestation, agriculture), erosion of mudbrick urban constructions (Rosen, 1986), and finally use of the basins as waste dumps. This underlines the importance of an explicit source-to-sink study integrating both the coastal area and the upland hinterland. Such high rates of harbor infilling were potentially detrimental to the medium- to long-term viability of harbor basins and impinged on the minimum 1 m draft depth. The harbor abandonment surface (HAS) This surface marks the “semi-abandonment” of the harbor basin. Recent studies have focused upon the role of natural hazards in explaining the decline or destruction of ancient Mediterranean harbors. While these factors may have had a role to play, it seems that the financial weight of maintaining harbor works in the face of the Mediterranean’s shifting political and economic makeup was simply too burdensome (Raban, 2009). A relative decline in harbor works after the late Roman and Byzantine periods is characterized by a return to “natural” sedimentary conditions comprising (1) coarse-grained sands and gravels in a coastal context and (2) terrestrial facies in fluvial environments. Following hundreds to thousands of years of artificial confinement, reconversion to a natural coastal parasequence is sometimes expressed by high-energy upper shoreface sands. This shoreline progradation significantly reduced the size of the basins, often landlocking the heart of the anchorages beneath thick tracts of coastal and fluvial sediments. Ancient harbor case studies: from natural to artificial ports Today, it is recognized that harbors should be studied within broader regional frameworks using a multidisciplinary methodology (Carayon, 2008; Blackman and Lentini, 2010). There is great variety in harbor types, and, broadly speaking, three areas or physical processes are important in influencing harbor location and design: (1) geographical situation, (2) site and local dynamics, and (3) navigation conditions dictated by the wind and wave climate. The diversity of contexts investigated during the past 20 years has brought to light some striking patterns. Numerous processes are important in explaining how these have come to be preserved in the geological record, including the distance from the present coastline, position relative to present sea level, and geomorphology (Marriner and Morhange, 2007). Ancient harbors can be divided into six non-exhaustive types on the basis of preservation. Sediment supply, human impacts, crustal changes, and coastal energy dynamics are significant in explaining how ancient harbors have been preserved in the geological record (Bony, 2013). Drowned harbors Drowned cities and harbors have long captured the public imagination and inspired research (Marinatos, 1960; Frost, 1963; Flemming, 1971; Bailey and Flemming, 2008), fueled by mediatized legends such as Atlantis (Collina-Girard, 2001; Gutscher, 2005) and the “biblical flooding” of the Black Sea (Yanko-Hombach et al., 2007a; Yanko-Hombach et al., 2007b; Ravilious, 2009; Buynevich et al., 2011). After the Last Glacial Maximum, when global sea level lay around 120 m below present, transgression of the continental platform gradually displaced coastal populations landwards until broad sea-level stability led to a sedentarization of populations along present coastlines (Van Andel 1989). The continental shelf between Haifa and Atlit (Israel) is one of the best-studied examples 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:09 Page Number: 9 Title Name: EOG HARBORS AND PORTS, ANCIENT 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 (Galili et al., 1988; Sivan et al., 2001). A series of submerged archaeological sites dating from the Pre-Pottery Neolithic B (8000 BP) and late Neolithic (6500 BP) were found at depths of 12 to 8 m and 5 to 0 m, attesting to the postglacial transgression of the Levantine coastline. Since 6000 cal BP, coastal site and port submersion can be attributed to crustal mobility (e.g., historical subsidence in eastern Crete and uplift on the western coast) and/or sediment failure in deltaic contexts. For example, on the western margin of the Nile Delta of Egypt, the coastal instability of the Alexandria area is responsible for a 5 m drowning of archaeological remains since antiquity (Empereur and Grimal, 1997; Goddio et al., 1998; Goiran, 2001; Fabre, 2004–2005). The subsidence has been variously attributed to seismic movements (Guidoboni et al., 1994) and Nile Delta sediment loading (Stanley et al., 2001; Stanley and Bernasconi, 2006). Approximately 22 km east of Alexandria, around Abu Qir bay, an 8 m collapse of the former Canopic lobe of the Nile is responsible for the drowning of two ancient seaport cities, Herakleion and East Canopus, during the eighth century AD (Tousson, 1922; Stanley et al., 2001; Stanley et al., 2004a; Stanley et al., 2004b). Italy’s Phlegraean Fields volcanic complex testifies to a very different crustal context that has led to a series of yo-yo land movements during the late Holocene. The ancient ports of Miseno, Baia, and Portus Julius are located inside a caldera (Gianfrotta, 1996; Scognamiglio, 1997; Figure 7). Since Roman times, tectono-volcanism inside this collapsed volcanic cone has led to significant shoreline mobility and is responsible for a 10 m submergence of the Roman harbor complexes (Dvorak and Mastrolorenzo, 1991). The pattern of movement inside the bay is spatially contrasted because around the fringes of the caldera the columns of the Roman market attest to an upper limit of marine bioerosion at 7 m above present sea level. Recent research suggests a series of post-Roman inflation-deflation cycles at both Pozzuoli (Morhange et al., 2006a) and Miseno (Cinque et al., 1991) linked to the interplay of deep magma inputs, fluid exsolution, and degassing (Todesco et al., 2004), all acting as drivers of rapid coastal change. Other studied examples of drowned cities include Helike and Kenchreai in the Gulf of Corinth, Greece (Kiskyras, 1988; Soter, 1998; Soter and Katsonopoulou, 1998; Rothaus et al., 2008) and Megisti on the island of Castellorizo, Greece (Pirazzoli, 1987). Uplifted harbors The best geoarchaeological evidence for uplifted harbors derives from the Hellenic arc, one of the most seismically active regions in the world (Stiros, 2005). In western Crete, Pirazzoli et al. (1992) have ascribed a 9 m uplift of Phalasarna harbor, founded in the fourth century BC, to high seismic activity in the eastern Mediterranean between the fourth to sixth centuries AD (Stiros, 2001). This episode is concurrent with a phase of Hellenic arc plate adjustment linked to uplift (1–2 m) in 9 Turkey, e.g., the uplifted harbor of Seleucia Pieria (Pirazzoli et al., 1991), Syria (Sanlaville et al., 1997), and parts of the Lebanese coastline (Pirazzoli, 2005; Morhange et al., 2006b). Phalasarna’s ancient harbor sediment record is of particular interest because its rapid uplift has possibly trapped tsunami deposits inside the basin (Dominey-Howes et al., 1998). The Gulf of Corinth constitutes a neotectonic graben separating the Peloponnese from mainland Greece (Moretti et al., 2003; Evelpidou et al., 2011). It is one of the most tectonically active and rapidly extending regions in the world (6–15 mm/year) with a marked regional contrast between its subsiding northern coast and an uplifting southern flank borne out by its geomorphological features and archaeology (Papadopoulos et al., 2000; Koukouvelas et al., 2001). Biological and archaeological proxies attest to pronounced spatial disparities in the amplitude of uplift. The position of the gulf’s ancient harbors can help to refine the recent tectonic history. The harbor of Heraion on the gulf’s northern coast is, for instance, modestly uplifted by around 1 m (Pirazzoli et al., 1994). The western harbor of Corinth at Lechaion is also uplifted. Emerged Balanus fossils indicating a former biological sea level 1.2 m above the basin surface have been dated to around 2470  45 BP, i.e., 375  120 cal BC (Stiros et al., 1996). The location of the port basin in a wellprotected depression suggests silting was already a problem during its excavation and not favorable to the basin’s long-term viability as a seaport (Morhange et al., 2012). At Aigeira, an artificial Roman harbor was functional between 100 AD and 250 AD (Papageorgiou et al., 1993). Biological and radiometric evidence from the city’s harbor structures attests to 4 m of uplift tentatively attributed to an earthquake around 250 AD (Stiros, 1998; Stiros, 2005). In a different geodynamic context, Holocene evolution of Etna’s coastline is associated with subduction of the African plate under the Eurasian plate. It presents a number of uplifted harbors, such as the neoria of the military harbor of Giardini-Naxos (Blackman and Lentini, 2010). This category of harbor is often poorly represented due to destruction by modern urbanization, e.g., the harbor of Kissamos, northwestern coast of Crete (Stefanakis, 2010). Landlocked harbors Around 6000 cal BP, the maximum marine ingression created an indented coastal morphology throughout the Mediterranean. During the ensuing millennia, these indented coastlines were gradually infilled by fluvial sediments reworked by longshore currents, culminating in a regularized coastal morphology. This process was particularly intense at deltaic margins. Coastal progradation as a driver of settlement and harbor changes is best represented by Ionia’s ancient ports in Turkey (Brückner, 1997), many of which are located inside infilled ria systems. Such rapid coastal change is 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:10 Page Number: 10 10 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 Title Name: EOG HARBORS AND PORTS, ANCIENT linked to two factors: (1) broad sea-level stability since 6000 cal BP; and (2) the morphology of these paleovalleys, which correspond to narrow, transgressed grabens with limited accommodation space (Kayan, 1996; Kayan, 1999). For example, the Menderes floodplain has prograded by 60 km during the past 7,000 years (Schröder and Bay, 1996). The best-studied examples include Troy (Kraft et al., 2003), where the harbor areas were landlocked by 2000 cal BP, and also Ephesus, Priene, and Miletos in Turkey (Brückner et al., 2005; Kraft et al., 2007). In Cyprus, Devillers (2008) has elucidated the infilling of the Gialia’s coastal embayment. The sedimentary archives attest to an easterly migration of the coastline. Human societies constantly adapted to this changing coastal environment as illustrated by the geographical shift of at least four ancient harbors: Early/Middle Bronze Age Kalopsidha, Middle/Late Bronze Age Enkomi, Graeco-Roman Salamina, and Medieval Famagusta. The latter is located on a rocky coast outside the paleo-ria. Despite the ecological attraction of estuaries and fluvial mouths for harbor location, ancient engineers were aware of the longer-term hazards to survival. Greek settlers, for instance, founded Marseille around 600 BC at the distal margin of the Rhone delta in order to avoid the problems of rapid siltation. It is only in instances of absolute necessity that artificial ports were located inside deltaic systems. The Imperial harbors of Portus on the Tiber delta are a classic example (Goiran et al., 2010). Eroded harbors Eroded harbors can result from two complementary geological processes: (1) a fall in sediment supply to the coastal zone and/or (2) the destruction of harbor works in areas exposed to high-energy coastal processes. The best examples of eroded harbors date from the Roman period, when natural low-energy roadsteads were no longer a prerequisite for harbor location. At many high- to medium-energy coastal sites across the Mediterranean, the Romans constructed large enveloping moles to accommodate mooring facilities and interface installations such as fishponds and industrial saltpans. Good examples of eroded ancient harbors include Carthage and the outer Roman basin of Caesarea Maritima (Raban, 2009). Fluvial harbors 1000 River harbors are not subject to the same geomorphologi1001 cal and sedimentary processes as coastal seaports, and 1002 therefore diagnostic harbor sediment signatures can be 1003 markedly different. Unfortunately, geoarchaeological 1004 study of such contexts has been relatively limited until 1005 now. It is nonetheless an interesting avenue for future 1006 research and provides opportunities with which to com1007 pare and contrast the coastal data (Milne and Hobley, 1008 1981; Good, 1991; de Izarra, 1993; Bravard and Magny, 1009 2002; Arnaud-Fassetta et al., 2003). In particular, current 1010 research has focused upon the relationships between 999 fluvial settlements, including their harbors, and flood hazards (Arnaud-Fassetta et al., 2003). The environmental challenges of fluvial harbors are linked to: (1) seasonal and exceptional flood episodes (Stewart and Morhange, 2009); (2) river mouth access and rapidly shifting longshore bar development; and (3) the lateral instability of riverbanks (Bruneton et al., 2001; Brown, 2008). The Egyptians and Mesopotamians were among the earliest western civilizations to engage in fluvial transportation, and primeval Bronze Age harbor works are known from the banks of the Nile at Memphis and Giza (Fabre, 2004–2005). Despite excavations at a number of sites on the Nile Delta, e.g., Tell El-Daba/Avaris and Tell el-Fara’in (Bietak, 1996; Shaw, 2000), the exact location of many of the river ports is equivocal. There has been extensive research looking at the Canopic branch of the Nile Delta coast (Stanley and Jorstad, 2006; Stanley, 2007). Geoarchaeological data show that the Ptolemaic and Roman city of Schedia (Egypt) once lay directly on the Canopic channel, which was active from the third to second centuries BC until the fifth century AD. Abandonment of the site resulted from the avulsion of Nile waters to the Bolbitic and later Rosetta branches in the east. The discovery of a series of active and abandoned channels around the Greek city of Naukratis (Egypt) attests to significant fluvial mobility during antiquity. These channels served as transport pathways for the ancient settlement, although the site’s fluvial port has never been precisely located (Villas, 1996). In the northeastern part of the Nile Delta, a number of sites on the now-defunct Pelusiac branch (Sneh and Weissbrod, 1973) have attracted geoarchaeological interest. Goodfriend and Stanley (1999) have shown that Pelusium, an important fortified city located at the mouth of the Pelusiac branch, was abandoned during the twelfth century AD following a large and rapid influx of Nile river sediment in the ninth century AD. This discharge in sediment led to the avulsion of a new distributory to the west, probably the Damietta branch. Aquileia in northeastern Italy is a well-studied example of a Roman fluvial harbor. A series of important waterways characterized the Aquileia deltaic plain during antiquity. These were channelized during the Roman period so as to ensure favorable conditions for navigation and to mitigate against the impact of floods (Arnaud-Fassetta et al., 2003). A similar evolution is attested at Minturnae (Italy), which controlled the bridge on the Appian Way over the Liris River. It occupied a prime location that allowed the Roman colony to evolve into a flourishing commercial center until its final abandonment around 590 AD. Recent geoarchaeological work undertaken at the mouth of the Tiber delta, around the ancient site of Ostia, has probed the evolution of the city’s ancient harbor, which serviced ancient Rome around 32 km upriver (Goiran et al., 2012). Problems of basin silting meant that the harbor had already experienced an important phase of sediment infilling by the first century AD (Goiran et al., 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:10 Page Number: 11 Title Name: EOG HARBORS AND PORTS, ANCIENT 2014). Continued late Holocene progradation dynamics have isolated ancient Ostia, which is now about 4 km from 1071 the present coastline. The silting of the harbor basin prob1072 ably acted as a precursor to the construction of Rome’s 1073 new port basin at Portus, although Ostia and the fluvial 1074 banks of the Tiber continued to accommodate smaller, 1075 shallow-draft vessels. 1076 At a number of sites, the excavation of ancient harbor 1077 quays has facilitated the precise reconstruction of fluvial 1078 bank mobility since antiquity. This can be linked to the 1079 vertical accretion of riverbanks by flooding and the grad1080 ual funneling of fluvial waters by human activities. In 1081 London, for instance, Milne (1985) has described 1082 a 100 m shift in the port’s waterfront between AD 1083 100 and today. Under a mesotidal fluvial regime, funnel1084 ing of the waterbody has led to a positive increase in tidal 1085 amplitude. A similar evolution is also attested at Bordeaux 1086 (France), where the staircasing of numerous quays and 1087 platforms has been described at two sites in the Garonne 1088 estuary (Gé et al., 2005). Three ancient and medieval plat1089 forms attest to a positive change in tidal amplitude of 1090 around 1.1 m during the twelfth to fourteenth centuries 1091 AD that can probably be linked to human impacts on the 1092 fluvial system. 1069 1070 Lagoonal harbors 1094 Since 6000 BP, spit accretion on clastic coasts has discon1095 nected a number of paleo-bays from the open sea. This 1096 process formed lagoons that have gradually infilled to 1097 yield rich geological archives. Lagoons offer natural pro1098 tection, and their use as anchorage havens has been wide1099 spread since early antiquity. Nevertheless, lagoons pose 1100 a number of challenges that explain why these contexts 1101 were largely avoided as harbors during later periods: 1102 (1) difficult accessibility, namely, the mobility of the outlet 1103 channel that was particularly problematic for navigation, 1104 and (2) seasonal fluctuations in lagoon level, especially 1105 in the case of large waterbodies at the margins of fluvial 1106 systems. 1107 Maryut lagoon lies at the northwestern margin of the 1108 Nile Delta, in a depression between two consolidated 1109 sandstone ridges of Pleistocene age (Flaux et al., 2011; 1110 Figure 8). The lagoon presently extends for 70 km on a 1111 SW-NE axis with a maximum width of 10 km. During 1112 antiquity, Nile inflow into the Maryut was supplied by 1113 the Canopic, the westernmost branch of the Nile. The 1114 Maryut’s location at the intersection between the Mediter1115 ranean Sea and a major fluvial system has driven impor1116 tant paleoenvironmental changes during the past 1117 8,000 years (Flaux, 2012; Flaux et al., 2012; Flaux et al., 1118 2013). It is also responsible for significant seasonal varia1119 tions in lagoon levels, driven by annual Nile flood cycles. 1120 There has been renewed interest in the Maryut because 1121 mounting archaeological evidence suggests that the 1122 lagoon was an important waterway during antiquity, with 1123 a densely occupied shoreline and numerous harbors and 1124 mooring sites (Blue and Khalil, 2010). Recent work by 1093 11 Flaux (2012) has demonstrated that the lagoon’s Hellenistic and Roman harbors present a steplike mooring architecture to accommodate these seasonal fluctuations. Similar annual variations of around 1.4 m are also attested in the Dead Sea and the Sea of Galilee (Hadas, 2011). Reinforced landing quays at the Roman harbor of Magdala (Israel) comprise a comparable architecture to offset such variation and avoid erosional undercutting (De Luca, 2009). Recent work has unearthed a wellpreserved harbor structure, extending for more than 100 m, which was functional during the Hellenistic and Roman periods (Sarti et al., 2013). Chronostratigraphic investigations have demonstrated that the harbor basin silted up and was abandoned during the Middle to Late Roman period (270–350 AD). Lagoonal systems were particularly conducive to endolagoonal harbor circulation. A number of lagoon strings were exploited in the Mediterranean during Roman times, most famously the Fossa Neronis (Italy) in the direction of Rome (Cuma, Campania), Narbonne in southern France (Sanchez and Jézégou, 2011), and the upper Adriatic lagoons between Istria and the Po (Degrassi, 1955). New archaeological data from the Maryut lagoon in Egypt also suggest that the basin possessed a series of harbor complexes and mooring sites during Hellenistic and Roman times (Blue and Khalil, 2010). At present, the archetype of a harbor lagoon is medieval Venice which operated very successfully as a port up until recent modification of its marginal marine system. Conclusions and future research directions The impact of ancient harbor geoarchaeology on our understanding of the archaeological record in waterfront areas is clear and explicit. We have presented methods for reconstructing ancient harbor landscapes at a wide range of temporal and spatial scales, drawing on geoscience techniques, paleoecology and archaeology. With particular emphasis on the Mediterranean region, we have concentrated on the description and illustration of selected case study examples drawn from different geomorphological contexts. These lay the foundations for more geographically extensive studies, integrating the archaeological record with sediment archives for many Holocene time periods. Some of the main advances made during the past 20 years include (1) the precise characterization of harbor facies in coastal contexts, using a variety of sedimentological, geochemical, and paleoecological proxies; (2) the characterization and intensity of human impacts in coastal areas (e.g., Véron et al., 2006); and (3) the scope to derive high-resolution RSL data (e.g., Morhange et al., 2001). Ancient harbor research is a rapidly evolving offshoot of geoarchaeology, and there is reason to be optimistic about its future prospects and applications. For the Mediterranean, as geographical gaps are gradually being filled and new research methods developed, more finite, regional- 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:11 Page Number: 12 12 Title Name: EOG HARBORS AND PORTS, ANCIENT scale interpretations are becoming possible at a variety of temporal scales. 1182 Current gaps in knowledge relate to the chronostra1183 tigraphic characterization of harbor facies in fluvial con1184 texts that, in the absence of archaeological structures, 1185 renders the precise localization of harbor basins particu1186 larly challenging. Furthermore, our understanding of 1187 ancient harbor geoarchaeology is biased towards later 1188 periods, particularly Greek and Roman ports. Major gaps 1189 remain with regard to the Bronze Age, and future studies 1190 must look to probe these earlier periods. While our under1191 standing of Mediterranean harbors continues to improve, 1192 it seems important to extend research to new geographical 1193 regions such as China, the Red Sea, and the Persian Gulf. 1194 One area of concern is the rise in catastrophic research in 1195 harbor contexts that mirrors the growth of neocatastrophic 1196 research during the past 20 years (Marriner et al., 2010; 1197 Marriner and Morhange, 2013). We advocate for the adop1198 tion of more nuanced approaches to the study of high1199 energy episodic events such as tsunamis and earthquakes. 1180 1181 1200 Bibliography 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 Abulafia, D., 2011. The Great Sea: A Human History of the Mediterranean. London: Allen Lane. Arnaud, P., 2005. Les routes de la navigation antique: Itinéraires en Méditerranée. Paris: Editions Errance. Arnaud-Fassetta, G., Carre, M.-B., Marocco, R., Maselli Scotti, F., Pugliese, N., Zaccaria, C., Bandelli, A., Bresson, V., Manzoni, G., Montenegro, M. E., Morhange, C., Pipan, M., Prizzon, A., and Siché, I., 2003. The site of Aquileia (northeastern Italy): example of fluvial geoarchaeology in a Mediterranean deltaic plain. Géomorphologie: relief, processus, environnement, 9(4), 227–245. Bailey, G. N., and Flemming, N. C., 2008. Archaeology of the continental shelf: marine resources, submerged landscapes and underwater archaeology. Quaternary Science Reviews, 27 (23–24), 2153–2165. Bard, K. A., and Fattovich, R., 2010. Recent excavations at the ancient harbor of Saww (Mersa/Wadi Gawasis) on the Red Sea. In D’Auria, S. H. (ed.), Offerings to the Discerning Eye. Brill: Leiden, pp. 33–38. Bellotti, P., Mattei, M., Tortora, P., and Valeri, P., 2009. Geoarchaeological investigations in the area of the imperial harbours of Rome. Méditerranée, 112, 51–58. Bicket, A. R., Rendell, H. M., Claridge, A., Rose, P., Andrews, J., and Brown, F. S. J., 2009. A multiscale geoarchaeological approach from the Laurentine shore (Castelporziano, Lazio, Italy). Géomorphologie: relief, processus, environnement, 2009(4), 241–256. Bietak, M., 1996. Avaris, the Capital of the Hyksos: Recent Excavations at Tell el-Dab’a. London: British Museum Press. Bini, M., Chelli, A., Durante, A. M., Gervasini, L., and Pappalardo, M., 2009. Geoarchaeological sea-level proxies from a silted up harbour: a case study of the Roman colony of Luni (northern Tyrrhenian Sea, Italy). Quaternary International, 206(1–2), 147–157. Blackman, D. J., 1982a. Ancient harbours in the Mediterranean, Part 1. International Journal of Nautical Archaeology and Underwater Exploration, 11(2), 79–104. Blackman, D. J., 1982b. Ancient harbours in the Mediterranean, Part 2. International Journal of Nautical Archaeology and Underwater Exploration, 11(3), 185–211. Blackman, D. J., and Lentini, M. C. (eds.), 2010. Ricoveri per navi militari nei porti del Mediterraneo antico e medievale. Bari: Edipuglia. Blue, L. K., and Khalil, E. (eds.), 2010. Lake Mareotis: Reconstructing the Past. Oxford: Archaeopress. British Archaeological Reports International Series, Vol. 2113. Boetto, G., 2012. Les épaves comme sources pour l’étude de la navigation et des routes commerciales: une approche méthodologique. In Keay, S. J. (ed.), Rome, Portus and the Mediterranean. Rome: British School at Rome. Archaeological Monographs, Vol. 21, pp. 153–173. Bony, G., 2013. Contraintes et potentialités naturelles de quelques sites portuaires antiques de Méditerranée et de Mer Noire (Fréjus, Ampurias, Kition, Istanbul, Orgame). PhD thesis, Aix-en-Provence, Aix-Marseille Université. Bony, G., Marriner, N., Morhange, C., Kaniewski, D., and Perinçek, D., 2012. A high-energy deposit in the Byzantine harbour of Yenikapı, Istanbul (Turkey). Quaternary International, 266, 117–130. Bottari, C., and Carveni, P., 2009. Archaeological and historiographical implications of recent uplift of the Peloro Peninsula, NE Sicily. Quaternary Research, 72(1), 38–46. Boyce, J. I., Reinhardt, E. G., and Goodman, B. N., 2009. Magnetic detection of ship ballast deposits and anchorage sites in King Herod’s Roman harbour, Caesarea Maritima, Israel. Journal of Archaeological Science, 36(7), 1516–1526. Braidwood, R. J., 1940. Report on two sondages on the coast of Syria, south of Tartous. Syria, 21(2), 183–226. Brandon, C., Hohlfelder, R. L., Oleson, J. P., and Stern, C., 2005. The Roman Maritime Concrete Study (ROMACONS): the harbour of Chersonisos in Crete and its Italian connection. Mé diterranée, 104, 25–29. Brandon, C., Hohlfelder, R. L., Oleson, J. P., and Rauh, N., 2010. Geology, materials, and the design of the Roman harbour of Soli-Pompeiopolis, Turkey: the ROMACONS field campaign of August 2009. International Journal of Nautical Archaeology, 39(2), 390–399. Braudel, F., 2002. The Mediterranean in the Ancient World. London: Penguin. Bravard, J.-P., and Magny, M. (eds.), 2002. Les fleuves ont une histoire: paléo-environnement des rivières et des lacs français depuis 15000 ans. Paris: Editions Errance. Brown, A. G., 2008. Geoarchaeology, the four dimensional (4D) fluvial matrix and climatic causality. Geomorphology, 101(1–2), 278–297. Brückner, H., 1997. Coastal changes in western Turkey; rapid delta progradation in historical times. In Briand, F., and Maldonado, A. (eds.), Transformations and Evolution of the Mediterranean Coastline. Monaco: Musée océanographique. Monaco, Bulletin de l’Institut océanographique, Special volume 18, pp. 63–74. Brückner, H., Müllenhoff, M., van der Borg, K., and Vött, A., 2004. Holocene coastal evolution of western Anatolia – the interplay between natural factors and human impact. In CIESM (Commission Internationale pour l’Exploration Scientifique de la mer Méditerranée) (ed.), Human Records of Recent Geological Evolution in the Mediterranean Basin – Historical and Archaeological Evidence (Santorini, Greece, 22–25 October 2003). Monaco: CIESM Workshop Monographs 24, pp. 51–56. Brückner, H., Vött, A., Schriever, M., and Handl, M., 2005. Holocene delta progradation in the eastern Mediterranean – case studies in their historical context. Méditerranée, 104, 95–106. Bruneton, H., Arnaud-Fassetta, G., Provansal, M., and Sistach, D., 2001. Geomorphological evidence for fluvial change during the Roman period in the lower Rhone valley (southern France). CATENA, 45(4), 287–312. 1241 1242 1243 1244 1245 1246 1247 1248 1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:12 Page Number: 13 Title Name: EOG HARBORS AND PORTS, ANCIENT 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338 1339 1340 1341 1342 1343 1344 1345 1346 1347 1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 Butzer, K. W., 1982. Archaeology as Human Ecology: Method and Theory for a Contextual Approach. Cambridge: Cambridge University Press. Butzer, K. W., 2005. Environmental history in the Mediterranean world: cross-disciplinary investigation of cause-and-effect for degradation and soil erosion. Journal of Archaeological Science, 32(12), 1773–1800. Butzer, K. W., 2008. Challenges for a cross-disciplinary geoarchaeology: the intersection between environmental history and geomorphology. Geomorphology, 101(1–2), 402–411. Butzer, K. W., Butzer, E., and Love, S., 2013. Urban geoarchaeology and environmental history at the Lost City of the Pyramids, Giza: synthesis and review. Journal of Archaeological Science, 40(8), 3340–3366. Buynevich, I. V., Yanko-Hombach, V., Gilbert, A. S., and Martin, R. E. (eds.), 2011. Geology and Geoarchaeology of the Black Sea Region: Beyond the Flood Hypothesis. Boulder: Geological Society of America Special Paper 473. Carayon, N., 2005. Le cothon ou port artificiel creusé. Essai de dé finition. Méditerranée, 104, 5–13. Carayon, N., 2008. Les ports phéniciens et puniques: gé omorphologie et infrastructures. PhD thesis, Strasbourg, Université Marc Bloch – Strasbourg II. Carsana, V., Febbraro, S., Giampaola, D., Guastaferro, C., Irollo, G., and Ruello, M. R., 2009. Evoluzione del paesaggio costiero tra Parthenope e Neapolis. Méditerranée, 112, 14–22. Casson, L., 1994. Travel in the Ancient World. Baltimore: Johns Hopkins University Press. Chen, Z., Zong, Y., Wang, Z., Wang, H., and Chen, J., 2008. Migration patterns of Neolithic settlements on the abandoned Yellow and Yangtze River deltas of China. Quaternary Research, 70(2), 301–314. Chittick, N., 1979. Early ports in the Horn of Africa. International Journal of Nautical Archaeology, 8(4), 273–277. Cinque, A., Russo, F., and Pagano, M., 1991. La successione dei terreni di età post-Romana delle terme di Miseno (Napoli): nuovi dati per la storia e la stratigrafia del bradisisma puteolano. Bolletino della Società Geologica Italiana, 110(2), 231–244. Coles, B., and Coles, J. M., 1989. People of the Wetlands: Bogs Bodies and Lake-dwellers. London: Thames and Hudson. Coles, J. M., and Lawson, A. J. (eds.), 1987. European Wetlands in Prehistory. Oxford: Clarendon. Collina-Girard, J., 2001. L’Atlantide devant le détroit de Gibraltar? Mythe et géologie. Comptes Rendus de l’Académie des Sciences. Sciences de la Terre et des planètes, 333(4), 233–240. de Izarra, F., 1993. Le fleuve et les hommes en Gaule romaine. Paris: Errance. De Luca, S., 2009. La città ellenistico-romana di Magdala/ Taricheae. Gli scavi del Magdala project 2007 e 2008: relazione preliminare e prospettive di indagine. Liber Annuus, 59, 343–562. Degrassi, A., 1955. I porti romani dell’Istria. In Fiocco, G. (ed.), Anthemon, Scritti di Archeologia e di Antichità Classiche in onore di Carlo Anti. Florence: G. C. Sansoni, pp. 119–169. Devillers, B., 2008. Holocene Morphogenesis and Anthropisation of a Semi-Arid Watershed, Gialias River, Cyprus. Oxford: Archaeopress. British Archaeological Reports International Series, Vol. 1775. Di Bella, L., Bellotti, P., Frezza, V., Bergamin, L., and Carboni, M. G., 2011. Benthic foraminiferal assemblages of the imperial harbor of Claudius (Rome): further paleoenvironmental and geoarcheological evidences. The Holocene, 21(8), 1245–1259. Dominey-Howes, D., Dawson, A., and Smith, D., 1998. Late Holocene coastal tectonics at Falasarna, western Crete: a sedimentary study. In Stewart, I. S., and Vita-Finzi, C. (eds.), Coastal Tectonics. London: Geological Society. Special Publication, Vol. 146, pp. 343–352. 13 Dvorak, J. J., and Mastrolorenzo, G., 1991. The Mechanisms of Recent Vertical Crustal Movements in Campi Flegrei Caldera, Southern Italy. Boulder: Geological Society of America. Special Paper, Vol. 263. Empereur, J.-Y., and Grimal, N., 1997. Les fouilles sous-marines du phare d’Alexandrie. Compte-rendus de l’Académie des Inscriptions et Belles-Lettres, 141(3), 693–717. Empereur, J.-Y., and Verlinden, C., 1987. The underwater excavations at the Ancient port of Amathus in Cyprus. International Journal of Nautical Archaeology, 16(1), 7–18. Evelpidou, N., Pirazzoli, P. A., Saliège, J.-F., and Vassilopoulos, A., 2011. Submerged notches and doline sediments as evidence for Holocene subsidence. Continental Shelf Research, 31(12), 1273–1281. Fabre, D., 2004–2005. Seafaring in Ancient Egypt. London: Periplus. Flaux, C., 2012. Holocene Palaeo-environments of the Maryut Lagoon in the NW Nile Delta, Egypt. PhD thesis, Aix-enProvence, Aix-Marseille Université. Flaux, C., Morhange, C., Marriner, N., and Rouchy, J.-M., 2011. Bilan hydrologique et biosédimentaire de la lagune du Maryût (delta du Nil, Egypte) entre 8 000 et 3 200 ans cal. B.P. Gé omorphologie, 2011(3), 261–278. Flaux, C., El-Assal, M., Marriner, N., Morhange, C., Rouchy, J.-M., Soulié-Märsche, I., and Torab, M., 2012. Environmental changes in the Maryut lagoon (northwestern Nile delta) during the last 2000 years. Journal of Archaeological Science, 39(12), 3493–3504. Flaux, C., Claude, C., Marriner, N., and Morhange, C., 2013. A 7500-year strontium isotope record from the northwestern Nile delta (Maryut lagoon, Egypt). Quaternary Science Reviews, 78, 22–33. Flemming, N. C., 1971. Cities in the Sea. Garden City, NY: Doubleday. Frost, H., 1963. Under the Mediterranean. London: Routledge and Kegan Paul. Frost, H., 1964. Rouad, ses récifs et mouillages. Prospection sousmarine. Annales Archéologiques de Syrie, 14, 67–74. Frost, H., 1973. The offshore island harbour at Sidon and other Phoenician sites in the light of new dating evidence. The International Journal of Nautical Archaeology and Underwater Exploration, 2(1), 75–94. Galili, E., Weinstein-Evron, M., and Ronen, A., 1988. Holocene sea-level changes based on submerged archaeological sites off the northern Carmel coast in Israel. Quaternary Research, 29(1), 36–42. Gambin, T., 2004. Islands of the Middle Sea: an archaeology of a coastline. In De Maria, L., and Turchetti, R. (eds.), Evolución paleoambiental de los puertos y fondeaderos antiguos en el Mediterráneo occidental. Soveria Mannelli: Rubbettino Editore, pp. 127–146. Gambin, T., 2005. The Maritime Landscapes of Malta from the Roman Period to the Middle Ages. PhD thesis, University of Bristol. Gaur, A. S., 2000. Harappan Maritime Legacies of Gujarat. New Delhi: Asian Publication Services. Gaur, A. S., and Vora, K. H., 1999. Ancient shorelines of Gujarat, India, during the Indus civilization (Late Mid-Holocene): a study based on archaeological evidences. Current Science, 77(1), 180–185. Gazda, E. K., 2001. Cosa’s contribution to the study of Roman hydraulic concrete: an historiographic commentary. In Goldman, N. W. (ed.), New Light from Ancient Cosa: Classical Mediterranean Studies in Honor of Cleo Rickman Fitch. New York: American Academy in Rome/Peter Lang, pp. 145–177. Gé, T., Migeon, W., and Szepertyski, B., 2005. L’élévation séculaire des berges antiques et médiévales de Bordeaux. Étude 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397 1398 1399 1400 1401 1402 1403 1404 1405 1406 1407 1408 1409 1410 1411 1412 1413 1414 1415 1416 1417 1418 1419 1420 1421 1422 1423 1424 1425 1426 1427 1428 1429 1430 1431 1432 1433 1434 1435 1436 1437 1438 1439 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:12 Page Number: 14 14 Au1 1440 1441 1442 1443 1444 1445 1446 1447 1448 1449 1450 1451 1452 1453 1454 1455 1456 1457 1458 1459 1460 1461 1462 1463 1464 1465 1466 1467 1468 1469 1470 1471 1472 1473 1474 1475 1476 1477 1478 1479 1480 1481 1482 1483 1484 1485 1486 1487 1488 1489 1490 1491 1492 1493 1494 1495 1496 1497 1498 1499 1500 1501 1502 1503 1504 1505 1506 Title Name: EOG HARBORS AND PORTS, ANCIENT géoarchéologique et dendrochronologique. Comptes Rendus Geoscience, 337(3), 297–303. Gébara, C., and Morhange, C., 2010. Fréjus (Forum Julii): le port antique/The Ancient Harbour. Journal of Roman Archaeology, Supplementary Series Number 77. Ghilardi, M., and Desruelles, S., 2009. Geoarchaeology: where human, social and earth sciences meet with technology. S.A.P.I. EN.S, 2(2): http://sapiens.revues.org/422. Giampaola, D., Carsana, V., and Boetto, G., 2004. Il mare torna a bagnare Neapolis. Parte II: dalla scoperta del porto al recupero dei relitti. L’Archeologo Subacqueo, 10(3), 15–19. Gianfrotta, P. A., 1996. Harbor structures of the Augustan Age in Italy. In Raban, A., and Holum, K. G. (eds.), Caesarea Maritima: A Retrospective After Two Millennia. Leiden: Brill Academic Publishers. Documenta et Monumenta Orientis Antiqui, Vol. 21, pp. 65–76. Giraudi, C., Tata, C., and Paroli, L., 2009. Late Holocene evolution of Tiber river delta and geoarchaeology of Claudius and Trajan Harbor, Rome. Geoarchaeology, 24(3), 371–382. Goddio, F., Bernand, A., Bernand, E., Darwish, I., Kiss, Z., and Yoyotte, J., 1998. Alexandria: The Submerged Royal Quarters. London: Periplus. Goiran, J.-P., 2001. Recherches géomorphologiques dans la region littorale d’Alexandrie, Egypte. PhD thesis, Aix-en-Provence, Université de Provence. Goiran, J. P., Tronchère, H., Salomon, F., Carbonel, P., Djerbi, H., and Ognard, C., 2010. Palaeoenvironmental reconstruction of the ancient harbors of Rome: Claudius and Trajan’s marine harbors on the Tiber delta. Quaternary International, 216(1), 3–13. Goiran, J.-P., Pavlopoulos, K. P., Fouache, E., Triantaphyllou, M., and Etienne, R., 2011. Piraeus, the ancient island of Athens; evidence from Holocene sediments and historical archives. Geology, 39(6), 531–534. Goiran, J.-P., Salomon, F., Pleuger, E., Vittori, C., Mazzini, I., Boetto, G., Arnaud, P., and Pellegrino, A., 2012. Résultats préliminaires de la première campagne de carottages dans le port antique d’Ostie. Chroniques des Mélanges de l’Ecole Française de Rome, 123(2), 2–7. Goiran, J.-P., Salomon, F., Mazzini, I., Bravard, J.-P., Pleuger, E., Vittori, C., Boetto, G., Christiansen, J., Arnaud, P., Pellegrino, A., Pepe, C., and Sadori, L., 2014. Geoarchaeology confirms location of the ancient harbour basin of Ostia (Italy). Journal of Archaeological Science, 41, 389–398. Good, G. L. (ed.), 1991. Waterfront Archaeology: Proceedings of the Third International Conference on Waterfront Archaeology held at Bristol, 23–26 September 1988. York: Council for British Archaeology. Goodfriend, G. A., and Stanley, J.-D., 1999. Rapid strand-plain accretion in the northeastern Nile Delta in the 9th century A. D. and the demise of the port of Pelusium. Geology, 27(2), 147–150. Guidoboni, E., Comastri, A., and Traina, G., 1994. Catalogue of Ancient Earthquakes in the Mediterranean Area up to the 10th Century. Roma: Istituto Nazionale di Geofisica. Gutscher, M.-A., 2005. Destruction of Atlantis by a great earthquake and tsunami? A geological analysis of the Spartel Bank hypothesis. Geology, 33(8), 685–688. Hadas, G., 2011. Dead Sea anchorages. Revue biblique, 118(2), 161–179. Haggi, A., 2006. Phoenician Atlit and its newly-excavated harbour: a reassessment. Tel Aviv: Journal of the Institute of Archaeology of Tel Aviv University, 33(1), 43–60. Haggi, A., and Artzy, M., 2007. The harbor of Atlit in northern Canaanite/Phoenician context. Near Eastern Archaeology, 70(2), 75–84. Halliday Saville, L., 1941. Ancient harbours. Antiquity, 15(59), 209–232. Hein, C. J., FitzGerald, D. M., Milne, G. A., Bard, K., and Fattovich, R., 2011. Evolution of a Pharaonic harbor on the Red Sea: implications for coastal response to changes in sea level and climate. Geology, 39(7), 687–690. Hesnard, A., 1994. Une nouvelle fouille du port de Marseille, Place Jules Verne. Compte-rendus de l’Académie des Inscriptions et Belles-Lettres, 138(1), 195–216. Hesnard, A., 1995. Les ports antiques de Marseille, Place JulesVerne. Journal of Roman Archaeology, 8, 65–77. Hesnard, A., 2004. Terre submergée, mer enterrée: une « géoarchéologie » du port antique de Marseille. In De Maria, L., and Turchetti, R. (eds.), Evolución paleoambiental de los puertos y fondeaderos antiguos en el Mediterráneo occidental. Soveria Mannelli: Rubbettino Editore, pp. 3–29. Hesse, A., 1998. Arguments pour une nouvelle hypothèse de localisation de l’Heptastade d’Alexandrie. In Empereur, J.-Y. (ed.), Études Alexandrines 1. Cairo: Institut Français d’Archéologie Orientale, pp. 21–33. Hesse, A., 2000. Archaeological prospection. In Ellis, L. (ed.), Archaeological Method and Theory: An Encyclopedia. New York: Garland Publishing, pp. 33–39. Heyvaert, V. M., and Baeteman, C., 2008. A middle to late Holocene avulsion history of the Euphrates river: a case study from Tell ed-Dēr, Iraq, Lower Mesopotamia. Quaternary Science Reviews, 27(25–26), 2401–2410. Hohlfelder, R. L., 1997. Building harbours in the early Byzantine era: the persistence of Roman technology. Byzantische Forschungen, 24, 367–380. Horden, P., and Purcell, N., 2000. The Corrupting Sea: A Study of Mediterranean History. Oxford: Blackwell Publishers. Ilves, K., 2009. Discovering harbours? Reflection on the state and development of landing site studies in the Baltic Sea region. Journal of Maritime Archaeology, 4, 149–163. Jondet, G., 1916. Les ports submergés de l’ancienne ı̂le de Pharos. Cairo: l’Institut égyptien. Mémoire 9. Kayan, I., 1996. Holocene coastal development and archaeology in Turkey. Zeitschrift für Geomorphologie Supplementband, 102, 37–59. Kayan, I., 1999. Holocene stratigraphy and geomorphological evolution of the Aegean coastal plains of Anatolia. Quaternary Science Reviews, 18(4–5), 541–548. Keay, S., and Paroli, L. (eds.), 2011. Portus and its Hinterland: Recent Archaeological Research. London: British School at Rome. Archaeological Monographs, Vol. 18. Keay, S., Millett, M., Paroli, L., and Strutt, K. (eds.), 2005. Portus: An Archaeological Survey of the Port of Imperial Rome. London: British School at Rome. Archaeological Monographs, Vol. 15. Keay, S., Earl, G., Hay, S., Kay, S., Ogden, J., and Strutt, K. D., 2009. The role of integrated geophysical survey methods in the assessment of archaeological landscapes: the case of Portus. Archaeological Prospection, 16(3), 154–166. Keller, F., 1866. The Lake-Dwellings of Switzerland and Other Parts of Europe. London: Longmans Green. Khalil, E., 2010. The sea, the river and the lake: all the waterways lead to Alexandria. Bollettino di Archeologia, volume speciale B/B7/5, 33–48. Kiskyras, D. A., 1988. The reasons for the disappearance of the ancient Greek town Helice (Eliki): geological contribution to the search for it. In Marinos, P. G., and Koukis, G. C. (eds.), Engineering Geology of Ancient Works, Monuments and Historical Sites. Rotterdam: Balkema, pp. 1301–1306. Koukouvelas, I. K., Stamatopoulos, L., Katsonopoulou, D., and Pavlides, S., 2001. A palaeoseismological and geoarchaeological investigation of the Eliki fault, Gulf of Corinth, Greece. Journal of Structural Geology, 23(2–3), 531–543. 1507 1508 1509 1510 1511 1512 1513 1514 1515 1516 1517 1518 1519 1520 1521 1522 1523 1524 1525 1526 1527 1528 1529 1530 1531 1532 1533 1534 1535 1536 1537 1538 1539 1540 1541 1542 1543 1544 1545 1546 1547 1548 1549 1550 1551 1552 1553 1554 1555 1556 1557 1558 1559 1560 1561 1562 1563 1564 1565 1566 1567 1568 1569 1570 1571 1572 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:12 Page Number: 15 Title Name: EOG HARBORS AND PORTS, ANCIENT 1573 1574 1575 1576 1577 1578 1579 1580 1581 1582 1583 1584 1585 1586 1587 1588 1589 1590 1591 1592 1593 1594 1595 1596 1597 1598 1599 1600 1601 1602 1603 1604 1605 1606 1607 1608 1609 1610 1611 1612 1613 1614 1615 1616 1617 1618 1619 1620 1621 1622 1623 1624 1625 1626 1627 1628 1629 1630 1631 1632 1633 1634 1635 1636 1637 1638 1639 Kraft, J. C., Rapp, G. R., Kayan, I., and Luce, J. V., 2003. Harbor areas at ancient Troy: sedimentology and geomorphology complement Homer’s Iliad. Geology, 31(2), 163–166. Kraft, J. C., Brückner, H., Kayan, I., and Engelmann, H., 2007. The geographies of ancient Ephesus and the Artemision in Anatolia. Geoarchaeology, 22(1), 121–149. Kramer, S. N., 1964. The Indus civilization and Dilmun: the Sumerian paradise land. Expedition, 6(3), 44–52. Laborel, J., and Laborel-Deguen, F., 1994. Biological indicators of relative sea-level variations and of co-seismic displacements in the Mediterranean region. Journal of Coastal Research, 10(2), 395–415. Lambeck, K., Anzidei, M., Antonioli, F., Benini, A., and Esposito, A., 2004. Sea level in Roman time in the Central Mediterranean and implications for recent change. Earth and Planetary Science Letters, 224(3–4), 563–575. Lauffray, J., Sauneron, S., and Traunecker, C., 1975. La tribune du quai de Karnak et sa favissa. Compte rendu des fouilles menées en 1971-1972 (2e campagne). Cahiers de Karnak, 5, 43–76. Lehmann-Hartleben, K., 1923. Die Antiken Hafenanlagen des Mittelmeeres: Beiträge zur Geschichte des Städtebaues im Altertum. Leipzig: Dietrich. Leung Tack, K. D., 1971–72. Étude d’un milieu pollué: le VieuxPort de Marseille. Influence des conditions physiques et chimiques sur la physionomie du peuplement de quai. Téthys, 3(4), 767–825. Leveau, P., 2005. L’archéologie du paysage et l’antiquité classique. Agri Centuriati, 2, 9–24. Leveau, P., Trément, F., Walsh, K., and Barker, G. (eds.), 1999. Environmental Reconstruction in Mediterranean Landscape Archaeology. Oxford: Oxbow Books. Marcus, E., 2002a. Early seafaring and maritime activity in the southern Levant from prehistory through the third millennium BCE. In Van den Brink, E. C. M., and Levy, T. E. (eds.), Egypt and the Levant: Interrelations from the 4th Through the Early 3rd Millennium BCE. London: Leicester University Press, pp. 403–417. Marcus, E., 2002b. The southern Levant and maritime trade during the Middle Bronze IIa period. In Oren, E., and Ahituv, S. (eds.), Aharon Kempinski Memorial Volume: Studies in Archaeology and Related Disciplines. Beer Sheva: Ben-Gurion University of the Negev Press. Studies by the Department of Bible and Ancient Near East, Vol. 15, pp. 241–263. Marinatos, S., 1960. Helice: a submerged town of Classical Greece. Archaeology, 13, 186–193. Markoe, G. E., 2000. Peoples of the Past: Phoenicians. Berkeley: University of California Press. Marriner, N., 2009. Géoarchéologie des ports antiques du Liban. Paris: L’Harmattan. Marriner, N., and Morhange, C., 2006a. Geoarchaeological evidence for dredging in Tyre’s ancient harbour, Levant. Quaternary Research, 65(1), 164–171. Marriner, N., and Morhange, C., 2006b. The ‘Ancient Harbour Parasequence’: anthropogenic forcing of the stratigraphic highstand record. Sedimentary Geology, 186(1–2), 13–17. Marriner, N., and Morhange, C., 2007. Geoscience of ancient Mediterranean harbours. Earth-Science Reviews, 80(3–4), 137–194. Marriner, N., and Morhange, C., 2013. Data mining the intellectual revival of “catastrophic” Mother Nature. Foundations of Science, 18(2), 245–257. Marriner, N., Morhange, C., Boudagher-Fadel, M., Bourcier, M., and Carbonel, P., 2005. Geoarchaeology of Tyre’s ancient northern harbour, Phoenicia. Journal of Archaeological Science, 32(9), 1302–1327. Marriner, N., Morhange, C., Doumet-Serhal, C., and Carbonel, P., 2006a. Geoscience rediscovers Phoenicia’s buried harbors. Geology, 34(1), 1–4. 15 Marriner, N., Morhange, C., and Doumet-Serhal, C., 2006b. Geoarchaeology of Sidon’s ancient harbours, Phoenicia. Journal of Archaeological Science, 33(11), 1514–1535. Marriner, N., Morhange, C., and Carayon, N., 2008. Ancient Tyre and its harbours: 5000 years of human-environment interactions. Journal of Archaeological Science, 35(5), 1281–1310. Marriner, N., Morhange, C., and Skrimshire, S., 2010. Geoscience meets the four horsemen? Tracking the rise of neocatastrophism. Global and Planetary Change, 74(1), 43–48. Martini, I. P., and Chesworth, W. (eds.), 2010. Landscapes and Societies. Dordrecht: Springer. Mason, O. K., 1993. The geoarchaeology of beach ridges and cheniers: studies of coastal evolution using archaeological data. Journal of Coastal Research, 9(1), 126–146. Mazzini, I., Faranda, C., Giardini, M., Giraudi, C., and Sadori, L., 2011. Late Holocene palaeoenvironmental evolution of the Roman harbour of Portus, Italy. Journal of Paleolimnology, 46(2), 243–256. Menotti, F., and O’Sullivan, A., 2012. The Oxford Handbook of Wetland Archaeology. Oxford: Oxford University Press. Milne, G., 1985. The Port of Roman London. London: Batsford. Milne, G., 2003. The Port of Medieval London. Stroud: Tempus. Milne, G., and Hobley, B. (eds.), 1981. Waterfront Archaeology in Britain and Northern Europe. London: Council for British Archaeology. Moretti, I., Sakellariou, D., Lykousis, V., and Micarelli, L., 2003. The Gulf of Corinth: an active half graben? Journal of Geodynamics, 36(1–2), 323–340. Morhange, C., and Marriner, N., 2010a. Mind the (stratigraphic) gap: Roman dredging in ancient Mediterranean harbours. Bollettino di Archeologia, volume speciale B/B7/4, 23–32. Morhange, C., and Marriner, N., 2010b. Palaeo-hazards in the coastal Mediterranean: a geoarchaeological approach. In Martini, I. P., and Chesworth, W. (eds.), Landscapes and Societies. Dordrecht: Springer, pp. 223–234. Morhange, C., Laborel, J., and Hesnard, A., 2001. Changes of relative sea level during the past 5000 years in the ancient harbor of Marseilles, Southern France. Palaeogeography Palaeoclimatology Palaeoecology, 166(3–4), 319–329. Morhange, C., Blanc, F., Schmitt-Mercury, S., Bourcier, M., Carbonel, P., Oberlin, C., Prone, A., Vivent, D., and Hesnard, A., 2003. Stratigraphy of late-Holocene deposits of the ancient harbour of Marseilles, southern France. The Holocene, 13(4), 593–604. Morhange, C., Hamdan Taha, M., Humbert, J.-B., and Marriner, N., 2005. Human settlement and coastal change in Gaza since the Bronze Age. Méditerranée, 104, 75–78. Morhange, C., Marriner, N., Laborel, J., Todesco, M., and Oberlin, C., 2006a. Rapid sea-level movements and noneruptive crustal deformations in the Phlegrean Fields caldera, Italy. Geology, 34(2), 93–96. Morhange, C., Pirazzoli, P. A., Marriner, N., Montaggioni, L. F., and Nammour, T., 2006b. Late Holocene relative sea-level changes in Lebanon, Eastern Mediterranean. Marine Geology, 230 (1–2), 99–114. Morhange, C., Pirazzoli, P. A., Evelpidou, N., and Marriner, N., 2012. Tectonic uplift and silting up of Lechaion, the western harbour of ancient Corinth, Greece. Geoarchaeology, 27(3), 278–283. Morhange, C., Marriner, N., Excoffon, P., Bonnet, S., Flaux, C., Zibrowius, H., Goiran, J.-P., and El Amouri, M., 2013. Relative sea-level changes during Roman times in the northwest Mediterranean: the 1st century A.D. fish tank of Forum Julii, Fréjus, France. Geoarchaeology, 28(4), 363–372. Négris, P., 1904a. Vestiges antiques submergés. Mitteilungen des Deutschen Archeologischen Instituts, Athenische Abteilung, 29, 340–363. 1640 1641 1642 1643 1644 1645 1646 1647 1648 1649 1650 1651 1652 1653 1654 1655 1656 1657 1658 1659 1660 1661 1662 1663 1664 1665 1666 1667 1668 1669 1670 1671 1672 1673 1674 1675 1676 1677 1678 1679 1680 1681 1682 1683 1684 1685 1686 1687 1688 1689 1690 1691 1692 1693 1694 1695 1696 1697 1698 1699 1700 1701 1702 1703 1704 1705 1706 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:13 Page Number: 16 16 1707 1708 1709 1710 1711 1712 1713 1714 1715 1716 1717 1718 1719 1720 1721 1722 1723 1724 1725 1726 1727 1728 1729 1730 1731 1732 1733 1734 1735 1736 1737 1738 1739 1740 1741 1742 1743 1744 1745 1746 1747 1748 1749 1750 1751 1752 1753 1754 1755 1756 1757 1758 1759 1760 1761 1762 1763 1764 1765 1766 1767 1768 1769 1770 1771 1772 Title Name: EOG HARBORS AND PORTS, ANCIENT Négris, P., 1904b. Nouvelles observations sur la dernière transgression de la Méditerranée. Comptes Rendus de l’Académie des Sciences, 2, 379–381. Nishimura, Y., 2001. Geophysical prospection in archaeology. In Brothwell, D. R., and Pollard, A. M. (eds.), Handbook of Archaeological Sciences. Chichester: Wiley, pp. 543–553. Oleson, J. P., 1988. The technology of Roman harbours. International Journal of Nautical Archaeology, 17(2), 147–157. Oleson, J. P., and Branton, G., 1992. The harbour of Caesarea Palaestinae: a case study of technology transfer in the Roman Empire. Mitteilungen, Leichtweiß-Institut für Wasserbau, 117, 387–421. Oleson, J. P., Brandon, C., Cramer, S. M., Cucitore, R., Gotti, E., and Hohlfelder, R. L., 2004. The ROMACONS Project: a contribution to the historical and engineering analysis of hydraulic concrete in Roman maritime structures. International Journal of Nautical Archaeology, 33(2), 199–229. Palley, R., 2010. Concrete: A Seven-Thousand-Year History. New York: Quantuck Lane Press. Papadopoulos, G., Vassilopoulou, A., and Plessa, A., 2000. A new catalogue of historical earthquakes in the Corinth rift, central Greece: 480 BC-AD 1910. In Papadopoulos, G. (ed.), Historical Earthquakes and Tsunamis in the Corinth Rift, Central Greece. Athens: National Observatory of Athens, Institute of Geodynamics, Vol. Publication 12, pp. 9–119. Papageorgiou, S., Arnold, M., Laborel, J., and Stiros, S. C., 1993. Seismic uplift of the harbour of ancient Aigeira, Central Greece. International Journal of Nautical Archaeology, 22(3), 275–281. Paris, J., 1915. Contributions à l’étude des ports antiques du monde grec. I. Note sur Léchaion. Bulletin de Correspondance Hellenique, 39, 5–16. Paris, J., 1916. Contributions à l’étude des ports antiques du monde grec. II. Les établissements maritimes de Délos. Bulletin de Correspondance Hellenique, 40, 5–73. Pirazzoli, P. A., 1987. Submerged remains of ancient Megisti in Castellorizo Island (Greece): a preliminary survey. International Journal of Nautical Archaeology, 16(1), 57–66. Pirazzoli, P. A., 2005. A review of possible eustatic, isostatic and tectonic contributions in eight late-Holocene relative sea-level histories from the Mediterranean area. Quaternary Science Reviews, 24(18–19), 1989–2001. Pirazzoli, P., and Thommeret, J., 1973. Une donnée nouvelle sur le niveau marin à Marseille à l’époque romaine. Comptes Rendus de l’Académie des Sciences, 277, 2125–2128. Pirazzoli, P. A., Laborel, J., Saliège, J. F., Erol, O., Kayan, I., and Person, A., 1991. Holocene raised shorelines on the Hatay coasts (Turkey): palaeoecological and tectonic implications. Marine Geology, 96(3–4), 295–311. Pirazzoli, P. A., Ausseil-Badie, J., Giresse, P., Hadjidaki, E., and Arnold, M., 1992. Historical environmental changes at Phalasarna harbor, west Crete. Geoarchaeology, 7(4), 371–392. Pirazzoli, P. A., Stiros, S. C., Arnold, M., Laborel, J., LaborelDeguen, F., and Papageorgiou, S., 1994. Episodic uplift deduced from Holocene shorelines in the Perachora Peninsula, Corinth area, Greece. Tectonophysics, 229(3–4), 201–209. Poidebard, A., 1939. Un grand port disparu, Tyr; Recherches aé riennes et sous-marines 1934–1936. Paris: Librairie Orientaliste Paul Geuthner. Poidebard, A., and Lauffray, J., 1951. Sidon, aménagements antiques du port de Saı̈da. Etude aérienne au sol et sous-marine, 1946–1950. Beyrouth: Imprimerie Catholique. Pomey, P., 1995. Les épaves grecques et romaines de la place JulesVerne à Marseille. Comptes Rendus de l’Académie des Inscriptions et Belles Lettres, 139(2), 459–484. Pomey, P., and Rieth, E., 2005. L’archéologie navale. Paris: Editions Errance. Pritchard, J. B., 1978. Recovering Sarepta, a Phoenician City. Princeton, NJ: Princeton University Press. Puglisi, D., 2001. Un arsenale marittimo l’Edificio T di Kommòs? Creta Antica, 2, 113–124. Purdy, B. A. (ed.), 1988. Wet Site Archaeology. Caldwell, NJ: The Telford Press. Raban, A., 1985. The Ancient Harbours of Israel in Biblical Times. In Raban, A. (ed.), Harbour Archaeology; Proceedings of the First International Workshop on Ancient Mediterranean Harbours. Caesarea Maritima, 24–28.6.83. British Archaeological Reports International Series 257. Oxford: British Archaeological Reports, pp. 11–44. Raban, A., 1987. Alternated river courses during the Bronze Age along the Israeli coastline. In Euzennat, M., Paskoff, R., and Trousset, P. (eds.), Déplacements des lignes de rivage en Mé diterranée d’après les données de l’archéologie. Paris: CNRS, pp. 173–199. Raban, A., 2009. The Harbour of Sebastos (Caesarea Maritima) in its Roman Mediterranean Context. Oxford: Archaeopress. British Archaeological Reports International Series 1930. Raban, A., and Holum, K. G. (eds.), 1996. Caesarea Maritima: A Retrospective after Two Millennia. Leiden: Brill Academic Publishers. Rao, S. R., 1979. Lothal, a Harappan Port Town, 2 vols. Memoir 78. New Delhi: Archaeological Survey of India. Rao, S. R. (ed.), 1988. Marine Archaeology of Indian Ocean Countries. Goa: National Institute of Oceanography in Dona Paula. Rao, S. R., 1991. Dawn and Devolution of Indus Civilization. New Delhi: Aditya Prakashan. Ravilious, K., 2009. ‘Biblical’ flood created present-day Mediterranean. The New Scientist, 204(2738), 12. Reinhardt, E. G., and Raban, A., 1999. Destruction of Herod the Great’s harbor at Caesarea Maritima, Israel – geoarchaeological evidence. Geology, 27(9), 811–814. Reinhardt, E. G., Patterson, R. T., and Schröder-Adams, C. J., 1994. Geoarchaeology of the ancient harbor site of Caesarea Maritima, Israel: evidence from sedimentology and paleoecology of benthic foraminifera. Journal of Foraminiferal Research, 24(1), 37–48. Reinhardt, E. G., Goodman, B. N., Boyce, J. I., Lopez, G., van Hengstum, P., Rink, W. J., Mart, Y., and Raban, A., 2006. The tsunami of 13 December A.D. 115 and the destruction of Herod the Great’s harbor at Caesarea Maritima, Israel. Geology, 34(12), 1061–1064. Rickman, G. E., 1988. The archaeology and history of Roman ports. International Journal of Nautical Archaeology, 17(3), 257–267. Rosen, A. M., 1986. Cities of Clay: The Geoarchaeology of Tells. Chicago: The University of Chicago Press. Rothaus, R., Reinhardt, E. G., and Noller, J. S., 2008. Earthquakes and subsidence at Kenchreai: using recent earthquakes to reconsider the archaeological and literary evidence. In Caraher, W. R., Hall, L. J., and Moore, R. S. (eds.), Archaeology and History in Medieval and Post-Medieval Greece: Studies on Method and Meaning in Honor of Timothy E. Gregory. London: Ashgate, pp. 53–66. Salomon, F., Delile, H., Goiran, J.-P., Bravard, J.-P., and Keay, S., 2012. The Canale di Comunicazione Traverso in Portus: the Roman sea harbour under river influence (Tiber delta, Italy). Géomorphologie : relief, processus, environnement, 2012(1), 75–90. Sanchez, C., and Jézégou, M.-P., 2011. Espaces littoraux et zones portuaires de Narbonne et sa région dans l’Antiquité. Monographies d’archéologie méditerranéenne 28. Lattes: l’Association pour le développement de l’archéologie en Languedoc-Roussillon. 1773 1774 1775 1776 1777 1778 1779 1780 1781 1782 1783 1784 1785 1786 1787 1788 1789 1790 1791 1792 1793 1794 1795 1796 1797 1798 1799 1800 1801 1802 1803 1804 1805 1806 1807 1808 1809 1810 1811 1812 1813 1814 1815 1816 1817 1818 1819 1820 1821 1822 1823 1824 1825 1826 1827 1828 1829 1830 1831 1832 1833 1834 1835 1836 1837 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:13 Page Number: 17 Title Name: EOG HARBORS AND PORTS, ANCIENT 1838 1839 1840 1841 1842 1843 1844 1845 1846 1847 1848 1849 1850 1851 1852 1853 1854 1855 1856 1857 1858 1859 1860 1861 1862 1863 1864 1865 1866 1867 1868 1869 1870 1871 1872 1873 1874 1875 1876 1877 1878 1879 1880 1881 1882 1883 1884 1885 1886 1887 1888 1889 1890 1891 1892 1893 1894 1895 1896 1897 1898 1899 1900 1901 Sanlaville, P., Dalongeville, R., Bernier, P., and Evin, J., 1997. The Syrian coast: a model of Holocene coastal evolution. Journal of Coastal Research, 13(2), 385–396. Sarti, G., Rossi, V., Amorosi, A., De Luca, S., Lena, A., Morhange, C., Ribolini, A., Sammartino, I., Bertoni, D., and Zanchetta, G., 2013. Magdala harbour sedimentation (Sea of Galilee, Israel), from natural to anthropogenic control. Quaternary International, 303, 120–131. Schröder, B., and Bay, B., 1996. Late Holocene rapid coastal change in Western Anatolia – Büyük Menderes Plain as a case study. Zeitschrift für Geomorphologie Supplementband, 102, 61–70. Scognamiglio, E., 1997. Aggiornamenti per la topografia di Baia sommersa. Archeologia Subacquea, 2, 35–45. Shaw, I., 2000. The Oxford History of Ancient Egypt. Oxford: Oxford University Press. Shaw, T., Sinclair, P., Andah, B., and Okpoko, A. (eds.), 1993. The Archaeology of Africa: Food, Metals and Towns. London: Routledge. Sivan, D., Wdowinski, S., Lambeck, K., Galili, E., and Raban, A., 2001. Holocene sea-level changes along the Mediterranean coast of Israel, based on archaeological observations and numerical model. Palaeogeography Palaeoclimatology Palaeoecology, 167(1), 101–117. Sneh, A., and Weissbrod, T., 1973. Nile Delta: the defunct Pelusiac branch identified. Science, 180(4081), 59–61. Soter, S., 1998. Holocene uplift and subsidence of the Helike Delta, Gulf of Corinth, Greece. In Stewart, I., and Vita-Finzi, C. (eds.), Coastal Tectonics. London: Geological Society Special Publication, Vol. 146, pp. 41–56. Soter, S., and Katsonopoulou, D., 1998. The search for ancient Helike, 1988–1995. Geological, sonar and bore hole studies. In Katsonopoulou, D., Soter, S., and Schilardi, D. (eds.), Ancient Helike and Aigialeia. Athens: The Helike Society Publications, pp. 67–116. Stanley, J.-D., 2007. Geoarchaeology: Underwater Archaeology in the Canopic Region in Egypt. Oxford: Oxford Center for Maritime Archaeology. Stanley, J.-D., and Bernasconi, M. P., 2006. Holocene depositional patterns and evolution in Alexandria’s eastern harbor, Egypt. Journal of Coastal Research, 22(2), 283–297. Stanley, J.-D., and Jorstad, T. F., 2006. Buried canopic channel identified near Egypt’s Nile Delta coast with radar (SRTM) imagery. Geoarchaeology, 21(5), 503–514. Stanley, D. J., and Warne, A. G., 1994. Worldwide initiation of Holocene marine deltas by deceleration of sea-level rise. Science, 265(5169), 228–231. Stanley, J.-D., Goddio, F., and Schnepp, G., 2001. Nile flooding sank two ancient cities. Nature, 412(6844), 293–294. Stanley, J.-D., Goddio, F., Jorstad, T. F., and Schnepp, G., 2004a. Submergence of ancient Greek cities off Egypt’s Nile delta – a cautionary tale. GSA Today, 14(1), 4–10. Stanley, J.-D., Warne, A. G., and Schnepp, G., 2004b. Geoarchaeological interpretation of the Canopic, largest of the relict Nile Delta distributaries, Egypt. Journal of Coastal Research, 20(3), 920–930. Stefanakis, M. I., 2010. Western Crete: from Captain Spratt to modern archaeoseismology. In Sintubin, M., Stewart, I. S., Niemi, T. M., and Altunel, E. (eds.), Ancient Earthquakes. Boulder: Geological Society of America. GSA Special Paper, Vol. 471, pp. 67–79. Stefaniuk, L., and Morhange, C., 2005. Évolution des paysages littoraux dans la dépression sud-ouest de Cumes depuis 4000 ans. La question du port antique. Méditerranée, 104(1–2), 49–59. 17 Stefaniuk, L., Morhange, C., Saghieh-Beydoun, M., Frost, H., Boudagher-Fadel, M. K., Bourcier, M., and Noujaim-Clark, G., 2005. Localisation et étude paléoenvironnementale des ports antiques de Byblos. Bulletin d’Archéologie et d’Architecture Libanaises, Hors-série 2, 19–41. Stewart, I. S., and Morhange, C., 2009. Coastal geomorphology and sea-level change. In Woodward, J. C. (ed.), The Physical Geography of the Mediterranean. Oxford: Oxford University Press, pp. 385–413. Stiros, S. C., 1998. Archaeological evidence for unusually rapid Holocene uplift rates in an active normal faulting terrain: Roman harbor of Aigeira, Gulf of Corinth, Greece. Geoarchaeology, 13(7), 731–741. Stiros, S. C., 2001. The AD 365 Crete earthquake and possible seismic clustering during the fourth to sixth centuries AD in the Eastern Mediterranean: a review of historical and archaeological data. Journal of Structural Geology, 23(2–3), 545–562. Stiros, S., 2005. Social and historical impacts of earthquake-related sea-level changes on ancient (prehistoric to Roman) coastal sites. Zeitschrift für Geomorphologie Supplementband, 137, 79–89. Stiros, S., Pirazzoli, P., Rothaus, R., Papageorgiou, S., Laborel, J., and Arnold, M., 1996. On the date of construction of Lechaion, western harbor of ancient Corinth, Greece. Geoarchaeology, 11(3), 251–263. Tallet, P., 2009. Les Égyptiens et le littoral de la mer Rouge à l’époque pharaonique. Comptes Rendus de l’Académie des Inscriptions et des Belles-Lettres, 2009, fasc. 2: 687–719. Tartaron, T. F., 2013. Maritime Networks in the Mycenaean World. Cambridge: Cambridge University Press. Testaguzza, O., 1970. Portus. Illustrazione dei porti di Claudio e Traiano e della città di Porto a Fiumicino. Roma: Julia Editrice. Todesco, M., Rutqvist, J., Chiodini, G., Pruess, K., and Oldenburg, C. M., 2004. Modeling of recent volcanic episodes at Phlegrean Fields (Italy): geochemical variations and ground deformation. Geothermics, 33(4), 531–547. Tousson, O., 1922. Mémoire sur les anciennes branches du Nil. Cairo: Institut Français d’Archéologie Orientale. Mémoires presentés à l’Institut d’Egypte, Vol. 4. Van Andel, T. H., 1989. Late Quaternary sea-level changes and archaeology. Antiquity, 63(241), 733–745. Van Beek, G., and Van Beek, O., 1981. Canaanite-Phoenician architecture: the development and distribution of two styles. EretzIsrael, 15, 70*–77*. Van de Noort, R., and O’Sullivan, A., 2006. Rethinking Wetland Archaeology. London: Duckworth. Véron, A., Goiran, J.-P., Morhange, C., Marriner, N., and Empereur, J.-Y., 2006. Pollutant lead reveals the pre-Hellenistic occupation and ancient growth of Alexandria, Egypt. Geophysical Research Letters, 33(6), L06409. Villas, C. A., 1996. Geological investigations. In Coulson, W. D. E. (ed.), Ancient Naukratis: Volume II, The Survey at Naukratis and Environs. Oxford: Oxbow Books. Oxbow Monograph, Vol. 60, pp. 163–175. Vött, A., Schriever, A., Handl, M., and Brückner, H., 2007. Holocene palaeogeographies of the central Acheloos River delta (NW Greece) in the vicinity of the ancient seaport Oiniadai. Geodinamica Acta, 20(4), 241–256. Walsh, K., 2004. Caring about sediments: the role of cultural geoarchaeology in Mediterranean landscapes. Journal of Mediterranean Archaeology, 17(2), 223–245. Walsh, K., 2008. Mediterranean landscape archaeology: marginality and the culture-nature ‘divide’. Landscape Research, 33(5), 547–564. 1902 1903 1904 1905 1906 1907 1908 1909 1910 1911 1912 1913 1914 1915 1916 1917 1918 1919 1920 1921 1922 1923 1924 1925 1926 1927 1928 1929 1930 1931 1932 1933 1934 1935 1936 1937 1938 1939 1940 1941 1942 1943 1944 1945 1946 1947 1948 1949 1950 1951 1952 1953 1954 1955 1956 1957 1958 1959 1960 1961 1962 1963 1964 1965 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:14 Page Number: 18 18 1966 1967 1968 1969 1970 1971 1972 1973 Title Name: EOG HARBORS AND PORTS, ANCIENT Yanko-Hombach, V., Gilbert, A. S., and Dolukhanov, P., 2007a. Controversy over the great flood hypotheses in the Black Sea in light of geological, paleontological, and archaeological evidence. Quaternary International, 167–168, 91–113. Yanko-Hombach, V., Gilbert, A. S., Panin, N., and Dolukhanov, P. M. (eds.), 2007b. The Black Sea Flood Question: Changes in Coastline, Climate and Human Settlement. Dordrecht: Springer. Zong, Y., Chen, Z., Innes, J. B., Chen, C., Wang, Z., and Wang, H., 2007. Fire and flood management of coastal swamp enabled first rice paddy cultivation in east China. Nature, 449(7161), 459–462. 1974 1975 1976 1977 Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:14 Page Number: 19 Title Name: EOG HARBORS AND PORTS, ANCIENT 19 ATLANTIC OCEAN Aquileia Bordeaux Frejus Marseille BLACK SEA Luni Rome Portus Minturnae Cuma  Pozzuoli Miseno Baia Portus Julius Byzantium/ Istanbul Coppa Nevigata Naples Troy Heraion Rachgoun Carthage Mahdia Giardini Naxos Piraeus Helike Aigeira Kenchreai 0 Seleucia Pierea Lechaion Kissamos Phalasarna MEDITERRANEAN SEA N Ephesus Priene Menderes ria 500 km Megisti Gialias Amathus Beirut Sidon Sarepta Dor Tyre Athlit Akko Caesarea Yavne Yam Herakleion East Wadi Ghazzeh Canopus Pelusium Naukratis Alexandria Tell Jezirat Fara’un Maryut Schedia Harbors and ports, ancient, Figure 1 Mediterranean harbor sites discussed in the text. Kition Bamboula Tabbat el-Hammam Byblos El-Daba Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:14 Page Number: 20 Title Name: EOG 20 HARBORS AND PORTS, ANCIENT Early Bronze 600 BC 575 BC Bargemon excavation site 500 BC Neolithic Bronze N 2nd -1st c. BC 1st -3rd c. AD Jules Verne excavation site 5th c. AD City hall Modern 0 20 m Vieux Port Harbors and ports, ancient, Figure 2 Coastal progradation in the ancient harbor of Marseille since Neolithic times. Chronostratigraphy and marine fauna fixed upon archaeological structures document a steady 1.5 m rise in relative sea level during the past 5,000 years. Sea level was broadly stable around the present datum between AD 1500 and the last century. Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:14 Page Number: 21 Title Name: EOG HARBORS AND PORTS, ANCIENT 0 120 m 21 1 km 80 m 40 m 0m -5m 0 15 knots - 10 m Anchoring - 30 m Docking ? Nahrel-Aou ali ? Zire Outer harbor Quay described in the Mission de Phénicie Present harbor Outcropping sandstone Saida Open harbor/ Crique "ronde" N Sidon-Dakerman Mont L ebano n M e d it e r ranean Sea - 20 m Wind rose Harbors and ports, ancient, Figure 3 Sidon’s ancient harbor areas (Adapted from Carayon (2008) and Marriner (2009)). Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:14 Page Number: 22 22 Title Name: EOG HARBORS AND PORTS, ANCIENT Harbors and ports, ancient, Figure 4 Harbor dredging in Naples (Photograph: D. Giampaola, Archaeological Superintendence of Naples). Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:15 Page Number: 23 Title Name: EOG HARBORS AND PORTS, ANCIENT Infill −1 Ostracod assemblages Texture 25 25 5075% 50 Palaeo -environmental interpretations A Semi-abandoned port B1 Byzantine harbour B2 Artificial Graeco-Roman harbour C Pocket beach / proto-harbour 75 % Coarse sands −2 Unit Mean sea level 0 1485 ± 30 BP 1420 -1300 cal. BP −3 Fine, silty sands 1910 ± 30 BP 1930 - 1770 cal. BP −4 2265 ± 30 BP 2350 - 2150 cal. BP −5 Fine sands 2360 ± 30 BP 2060 - 1890 cal. BP 2245 ± 35 BP 2350 - 2150 cal. BP −6 Coarse sands 5730 ± 30 BP 6260 - 6020 cal. BP −7 6400 ± 35 BP 7760 ± 40 BP 7780 ± 40 BP D Shelly silts and clays 1 Plastic clays substratum Sand N 1000 10000 Number of ostracods per 10 g of sand (log scale) Lagoonal Coastal Marine Marine lagoonal Phoenician sea wall m ar se gr av el s e ed 100 Gravels Sands Silts & clays co fin m cl ay si lt −9 10 5 7800 ± 40 BP 8350 - 8160 cal. BP Lagoon T5 T1 T4 T2 T9 T6 Tyre Bronze Age coastline 0 −8 m Depth (m) 23 Buried harbor basin Ancient caus eway 0 500 m Tombolo 10 m Submerged urban quarters 5m 0m Harbors and ports, ancient, Figure 5 Chronostratigraphic evolution of Tyre’s ancient northern harbor since the Bronze Age (core T9). Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:15 Page Number: 24 Title Name: EOG 24 HARBORS AND PORTS, ANCIENT Ancient Harbor Parasequence Log Lithofacies association Key stratigraphic surfaces Iron Age Ancient harbor muds Bronze Age proto-harbors Transition phase Upper shoreface Roman & Byzantine engineering apogee ~ ~ 3 Harbor Abandonment Surface 2 Harbor Foundation Surface 1 Maximum Flooding Surface Lower/ middle shoreface ~ ~ cl ay si lt fin e m e co d ar se Flooding surface Sand Prograding upper shoreface/foreshore (harbor abandonment) Ancient harbor muds Lower/middle shoreface sands Harbor dredging Key stratigraphic surfaces Harbors and ports, ancient, Figure 6 Chronostratigraphic evolution of ancient Mediterranean harbors in coastal areas. Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:15 Page Number: 25 Title Name: EOG HARBORS AND PORTS, ANCIENT 25 Harbors and ports, ancient, Figure 7 Pozzuoli’s drowned harbor remains presently 10 m below mean sea level. The site lies inside a caldera, where shoreline mobility is attributed to volcanism and faulting (Photograph: Centre Jean Bérard, Naples). Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:16 Page Number: 26 26 Title Name: EOG HARBORS AND PORTS, ANCIENT Harbors and ports, ancient, Figure 8 Evolution of the Maryut lagoon during the past 3,000 years (From Flaux, 2012). The general aridification trend described during this period appears to be linked to the gradual decline of the Canopic branch of the Nile, which supplied the Maryut lagoon with freshwater. Comp. by: D.Prabhakharan Stage: Galleys Chapter No.: 119 Date:21/7/14 Time:22:29:17 Page Number: 27 Title Name: EOG Author Query Form Encyclopedia of Geoarchaeology Chapter No: 119 _________________________________________________________________________________ Query Refs. Details Required AU1 Please provide volume and page number for Gébara and Morhange (2010) if applicable. Author’s response