Download MMHE_RF_Roman_mw3[1]

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Wales in the Early Middle Ages wikipedia , lookup

Migration Period wikipedia , lookup

Transcript
Developing a ‘Maritime and Marine Historic Environment Research Framework’ for England
ROMAN section (10,000 words max)
Executive Summary
Introduction
Caesar's ‘Gallic Wars’ provides the setting against which cross-Channel contacts, both before and after
conquest, can be measured thus offering a unique insight into the process of Romanization, conquest,
occupation and ultimately withdrawal. The Channel was a barrier which had to be crossed, so it has a special
resonance for maritime archaeology; there must be more direct evidence for maritime contacts in the form of
potential wrecks located off the southern coast of Britain (Muckelroy 1978: 143)
This is a research framework for England but will include best available data from Guernsey, Wales and
northern France.
Resource Assessment
1. Coastal Evolution
Coastal change over the last two millennia has produced a modern coastline significantly different from that of
the Roman period. The natural processes involved include erosion, (with landslips resulting from instability in
the underlying geology on some coasts) and accretion, both driven by a range of factors, but most significantly
by changes in climate, relative sea level (RSL) and sediment supply.
In places there has been sudden catastrophic breaching of natural dune and shingle beach barriers that
formerly protected low-lying areas. Anthropogenic land-claim has further modified the coast, and this most
probably began during the Roman period, though there is much dispute about this. Textbook maps of Roman
Britain based on modern geography may therefore be very misleading: much of the coastline of Britannia has
been entirely lost, whilst other parts of it are deeply buried beneath later sediments.
Change in RSL, (a term which encompasses both global (eustatic) change in sea level and regional (isostatic)
change in the elevation of land surfaces, as well as local effects such as land subsidence due to compression of
soft sediments) is conventionally reconstructed using Sea Level Index Points (ie Long and Roberts 1997).
Typically these comprise contacts between sediment units, dated by radiocarbon. However, calibrated
radiocarbon dates may not provide the refinement required for the reconstruction of sea-level change within
the Roman period.
Additional information may be gleaned from the elevation of Roman quayside surfaces, although this too is
problematic, unless there is good palaeoecological evidence to establish the position of quays in relation to the
tidal range. Moreover, the top surface of a quay only demonstrates a probable RSL at a given point, which
cannot be extrapolated elsewhere (Toft 1992). However, in terms of our understanding of the Roman coast
though clearly significant, RSL matters less than coastline morphology.
Rates of coastal cliff erosion are related largely to geology, wave climate and the incidence and severity of
storms. At present the most rapidly eroding cliffs in England are on the coasts of east Yorkshire and East
Anglia (where the geology largely comprises soft unconsolidated glacial till and sands), and parts of south-east
England (where deposits with similar mechanical properties but of earlier date outcrop). Land-slips are
prevalent in parts of Dorset and the Isle of Wight on shales and clays; elsewhere there are local ‘hot-spots’ of
erosion of unconsolidated cliffs (McInnes 2009). Some records exist of the relatively recent destruction by
erosion of some Roman coastal sites; the destruction of the shore fort of Walton Castle in Suffolk is depicted
on an engraving dated 1786. Other sites are likely to have disappeared far earlier.
Several of the small forts on the Yorkshire coast (at Huntcliff, Goldsborough, Ravenscar, Scarborough and
Filey), often described as ‘signal stations’ (Wilson 1989), are sited on modern cliff-edges and have been only
partly excavated before being lost. Indeed, given the massive coastal erosion that has occurred since the late
Roman period along the north-east, Holderness, Lincolnshire and East Anglian coasts it is perfectly possible
that this whole defensive system originally extended much further south and north, but has been lost to
erosion (Petts & Gerrard 2006).
Assessing the extent of land loss is problematic. In some places historical accounts and map regression studies
permit some reconstruction of long-term erosion rates. At Reculver, for example, the loss of about half of the
fort can be reconstructed in some detail (Philp 1996). Elsewhere modern trends can be used. Along the north
and north-east coasts of Norfolk, change since the Roman period has been dominated by erosion and
landward movement of barrier beaches and dune systems. By extrapolating the present rate of barrier beach
movement (about 1m per year), a Roman coast in the order of 2km further seawards is proposed; rates of cliff
erosion are also around 1-2m per year in that area (Murphy 2005). However, this pre-supposes a mean rate of
erosion annually comparable to current rates.
Besides the fact that erosion is episodic, whereby unconsolidated cliffs erode in a cyclical manner, climate and
consequently wave climate has not been constant. Storm incidence and severity at any specific location in the
past is difficult to reconstruct from scientific evidence, although analysis of historical reports from the
medieval period indicates an exceptional phase of severe and sustained storms in the late thirteenth to
fourteenth centuries (Rippon 2000: 30-31), with concomitant enhanced rates of erosion.
It is far more straightforward to reconstruct Roman coastlines in low-lying accreting areas of coast, where the
preserved sedimentary sequence and architecture can be investigated by means of boreholes,
palaeoecological analysis and scientific dating. Some areas, like the Thames Estuary, the East Anglian Fens and
Romney Marsh, have been studied in detail (Devoy 1979; Waller 1994; Rippon 2000, 190-195); others not at
all. On such coastlines, although coastal change is ultimately driven by climate and RSL change, it has been
largely determined by much more regional or even local environmental factors over this period.
Reconstructed events in one embayment or estuary cannot be assumed to apply in other places. It is not
possible here to discuss more than a few studies, and the palaeogeographic reconstructions that can be
derived from them.
In the Fens at this time there was settlement expansion onto the western part of a zone of estuarine/marine
silts (the Terrington Beds), lying between the peat fen to the south and the estuaries of the Ouse and Nene to
the north. However, a late third-fourth century saltern at Middleton demonstrates that the lower Nar was
certainly tidal at this time.
Moreover, biological analysis of sediments related to the Fen Causeway at Nordelph demonstrates that the
Roman road crossed salt marsh and suffered catastrophic marine flooding at least once during its lifetime,
before being ultimately overwhelmed by laminated marine silts (Crowson et al 2000). In Broadland
destruction of an earlier coastal barrier on the site of Great Yarmouth permitted development of a major
Bure/Yare/Waveney estuary and fully estuarine conditions extended to within 7km of Caistor (Venta
Icenorum) by the late Roman period (Coles and Funnell 1981). Access to the estuary was controlled by two
forts, at Caister-on-Sea and Burgh Castle. To the north lay an island later known as Flegg, isolated from the
mainland by the Bure and by an estuarine channel out to the North Sea, of unknown width, north of
Winterton.
Although the outer Thames estuary had essentially adopted its present form by the Roman period it seems
probable that the main approach from the near Continent was along the Wantsum Channel to the south of the
Isle of Thanet. This seaway provided a sheltered route to London, which avoided the hazardous areas of the
Goodwin Sands and North Foreland. Coastal forts were constructed at its southern and northen ends, at
Richborough and Reculver respectively. Historical sources demonstrate that it remained navigable until the
Middle Ages (Lydden Valley Research Group 2006).
Further upstream, the prehistoric river channel at London was originally approximately three times as wide as
it is today, but on the Southwark side there were several substantial ‘eyots’ (stable sand islands), which were
inhabited in the Middle Bronze Age. Diatom analysis of sediments indicate that by about 1200cal BC tidal
waters had extended upstream of Westminster, and the eyots were abandoned. By the time of the Roman
invasion the tidal head had migrated downstream, the eyots were once more habitable when revetted, and
could linked together to form a bridging point to the north bank (Sidell et al. 2000).
Sea walls almost certainly of Roman date are known or suspected from the Solway Firth, East Anglian
Fenlands, East Kent, Somerset and Severn Estuary (Allen & Fulford 1990; Fulford et al 1994; Hall & Coles 1994;
Lydden Valley Research Group 2006; Rippon 1992; Simmons 1980). Land claim during the Roman period in the
Severn Estuary has been inferred from several lines of evidence, in particular the surface elevation of
reclaimed land, and the presence of surface scatters of Roman pottery, which imply settlement and/or
manuring of fields with domestic and agricultural waste. Documentary references, and the survival of
medieval ridge-and-furrow may also indicate a medieval or earlier date (Allen & Fulford 1990).
In the upper Severn, the Great Wall of Elmore runs for 800m across the alluvium at Bridgemacote, with a stone
revetment along its south west side. This suggests that it was a sea defence rather than a flood defence for
previously reclaimed land, so it could be one of the finest examples of a Roman sea defence to be identified in
Britain. However, there is no direct archaeological evidence for this structure, and its early date is questioned
(Allen & Fulford 1990).
Given the undoubted ability of Roman engineers to undertake reclamation projects, the rather tenuous and
debatable evidence for Roman land-claim in England seems surprising. The most likely explanation is that
there was no land-hunger and so no incentive to undertake such costly projects: although large areas of
Roman Britain were under cultivation, areas of uncleared woodland that could have been converted to
farmland more easily, still survived.
The abandonment of English and near Continent coastal marshes between the third and fifth centuries AD
probably resulted from a range of factors including marine transgression, economic change, political
insecurity, and large-scale population movements (Rippon 2000: 138-151).
2. Maritime settlement and exploitation
Harbours
It is generally accepted that, with the notable exception of London, comparatively few remains of Roman
harbours and quays have been identified in Britain. Large numbers of harbours must have existed as an island
province like Britain was heavily dependent on its sea communications with the Continent. Their absence in
the archaeological record must in part reflect the vulnerability of harbour installations to destruction as a
result of coastal or river change, or as a result of continued use of natural harbour sites down to the present
day (Jones and Mattingly 1990: 198). For example, despite the riverside setting of Roman Gloucester efforts to
locate the Roman and early medieval waterfronts have failed (Hurst 1999: 123). Similarly, despite
considerable efforts to locate quays at York and Lincoln for which there is epigraphic evidence for overseas
trade, their whereabouts remain unknown (Ottaway 1993, 85). What of all those Romano-British coastal sites
like Harwich, Dover, Clausentum, Hamworthy or Radipole where on-shore Roman settlement is known?
However, there is sufficient archaeological evidence to demonstrate the scale and extent of harbour, port and
landing place development in Roman Britain. The following representative sample, chosen for its diversity, has
been grouped under the general headings of ‘coastal’ denoting harbours with direct access to the open sea,
‘estuarine’ denoting sites located on or near an estuary, and ‘inland’ denoting riverine locations, sometimes
many kilometres from the coast.
Coastal
At the closest point to continental Europe, Dover, succeeded Richborough as the principal cross-Channel port
for both military and mercantile traffic. A pharos (lighthouse) was built on each of the headlands overlooking
the harbour, and the remains of that on the Eastern Heights still display some 13 m of Roman construction.
The remains of a massive breakwater that formed a protected anchorage at the mouth of the River Dour were
found some 230m to the east of the fort. A probable quay and timber jetty were found to the west of the
inlet, while a structure closer to the mouth of the inlet has been interpreted as a continuation of the
harbourside. The headquarters for the British squadron of the classis Britannica was commenced in AD 116,
but after several phases of refurbishment and re-occupation up to the early third century it was then
abandoned and was probably demolished (Philp 1981).
The fort of Arbeia, situated at the mouth of the river Tyne, is the most extensively excavated Roman military
supply-base in the Roman Empire, yet no port facilities have been found. The original Hadrianic port was
converted into a supply base, with fifteen granaries, to support the AD 208–11 Scottish campaigns of
Septimius Severus. It is suggested that the Severan campaigns were essentially punitive in nature, employing
water transport on a massive scale to efficiently apply and sustain crushing military force at any chosen point
(Martin 1992, 20–21; 25–29).
A further nine granaries were added between 222 and 235 when Arbeia functioned as the major port
supporting the garrisons of the eastern sector of Hadrian’s Wall. The Notitia Dignitatum records that the
Numerus Barcariorum Tigrisiensum (Tigris Bargemen) were stationed at Arbeia during the late fourth century.
It has been assumed that they were engaged in transportation and lighterage but it is more likely that the unit
functioned as a maritime coastal defence force against seaborne raiders.
Plymouth Sound provides an outstanding natural harbour; the probability that it served as a significant Roman
port/settlement is supported both by the use of Mount Batten promontory as an entrepôt in the Iron Age
(Cunliffe 1988) and by the coin evidence. At least eight hoards and over fifty separate coin finds, dating from
the first to fourth centuries, have been recovered in the vicinity. In addition, significant quantities of Roman
building materials, including tiles, have been recovered from the foreshore of Sutton Pool, and other Roman
material has been found at various locations on the banks of both the rivers Tavy and Tamar.
Estuarine
The most extensive Romano-British waterfront yet discovered is London where excavations have revealed
considerable details of the development of the port from the late first century through the mid third century,
until its decline in the fourth century. Each successive quay was laid further into the river level than its
predecessor, indicating that the tidal level of the Thames fell by as much as 1.5m (Milne 1995: 78–81; 1985,
22–33; Brigham 2001: 15–49). A workable depth of water was maintained by constructing the bases and tops
of successive quays at lower levels than their predecessors. However, London’s modern pre-eminence as
paramount Roman port may not reflect its importance in antiquity (Milne 1985: 147; contra Morris 1982: 162)
Excavations on the river Dee at Chester in 1885 revealed an ancient riverbed at a depth of about 6m below
present ground level. Roman material was found, including bricks, tiles, samian and other pottery types, and
a lead ingot bearing a date of manufacture of AD 74. Lengths of oak timber were found, some 3m in length
and 0.3m in diameter, with the lower ends encased in iron sheaths and embedded in concrete (Shrubsole
1887: 80). This technique was commonly used in the construction of bridges, wharves, and jetties during the
Roman period so this may represent the remains of a landing stage projecting into the deepest part of the
river channel (Mason 2002: 64–72).
At Sudbrook on the northern coast of the Severn estuary, a Roman garrison was installed within the ramparts
of an Iron Age promontory fort, presumably to control a ferry crossing from the English side to Portskewett. A
road runs to the main Caerwent road at Crick and probably accords with a decision to ship supplies across the
Severn estuary, rather than by the long and circuitous route by road via Gloucester. Coins recovered from the
foreshore at Black Rock span 300 years of Roman occupation, from Claudius to Gratian and indicate dedicatory
offerings after a safe crossing.
Riverine
The colonia at Lincoln lay on the navigable river Witham, some 60km from its entry into the Wash and it is
probable that the course of the river was subject to some Roman modification. From Lincoln, the Fosse Dyke
ran in the direction of Littleborough and the Car Dyke ran south towards Water Newton and possibly as far as
Cambridge, and connected with the navigable rivers Welland and Trent. Whether the canals were used for
navigation or just for drainage is still a matter of debate. There has been significant progress in establishing the
location of the Roman waterfront (up to 100m distant from the present position), but other than a 6m stretch
of stone wall no significant features have been found (Jones 2002: 107).
However, the process of land reclamation included the dumping of waste material; painstaking excavation has
led to the recovery of the remains of table and drinking vessels (including a high proportion of samian ware),
jewellery, glass, coins, fragments of weapons, and armour and even leather sandals. A wooden writing tablet
and nineteen styli, together with a copper-alloy balance might also be considered to be evidence of waterfront
commercial activity, while overseas trade is indicated by the dedicatory inscription of M. Aurelius Lunaris, a
wine merchant from Bordeaux (ibid. Figs. 25, 64, 10).
The ‘small town’ of Worcester was situated on a major road and on the banks of a major navigable river (the
Severn) with easy access to the Western seaways and was therefore able to take advantage of an integrated
transport system. This was crucial as it was engaged in the metal industry on a substantial scale, making use of
processed ore from the Forest of Dean. The river was probably fordable; the existence of a Roman bridge over
the river has yet to be proved, although it seems a reasonable proposition and the tactical importance of the
river crossing suggests an early fort. Two wooden water pipes, with surviving iron junctions, suggest the
existence of an aqueduct and distribution system. The absence of late-period coins suggests abandonment in
the mid third or early fourth century.
An iron foundry, with at least six smelting hearths, was established in the third century; timber buildings east
of the present Cathedral ran along a street made entirely of iron slag (Burnham & Wacher 1990: 232–4). Ore
was brought up-river from the Forest of Dean from, for example, the villa at Woolaston on the banks of the
River Severn, where excavations revealed evidence of a highly organised enterprise for the production of iron
blooms, covering an area of about 7,250 square metres. The working of the blooms into billets probably took
place near to Ley Pill and the iron was shipped to Worcester in billet form, rather than as finished articles
(Fulford and Allen 1992: 159–215). It is also possible that lead from mines near Felindre, some 30km above
Leintwardine, and iron ore from mines above Ludlow was carried down to forges in Worcester in barges.
The settlement site at Heronbridge stands on the west bank of the River Dee 2km south of Chester., An
existing streambed was deepened in the second century to enable construction of a ramp down to the edge of
an inlet from the Dee, which had been artificially straightened. There were several rock-cut pits and other
features that were probably mooring posts to enable barges to be tied-up prior to loading or unloading.
Classis Britannica
Although the classis Britannica is not mentioned by contemporary historians, there is firm evidence for its
existence from c AD100 to cAD250, and seven fleet prefects are known from inscriptions dated c AD136 - 208
(Spaul 2002 47). For example, an inscription found in Camerinum, Umbria dated to AD 138-61 recorded that
Marcus Maenius Agrippa held the posts both of Prefect of the classis Britannica and Procurator of the province
of Britannia, a combination of offices which, it has been suggested, implies that the classis Britannica operated
as a branch of the Procurators office, effectively functioning as the state haulage company (Milne 2007).
In addition, there are a series of undated tile stamps, principally from Boulogne and Desvres in France, Dover,
Beauport Park and associated sites in the Weald of Kent and Sussex as well as London, and three building
inscriptions from Hadrian’s Wall. These inscriptions identify army units involved in wall building suggesting
that, as with other naval units, the classis Britannica was organised along the same lines as any other unit of
milites undertaking routine military functions, as well as operating ships and presumably their associated
infrastructure such as docks and ship yards. The classis Britannica seems to have acted more as an army
service corps supporting the government and provincial military than as a navy in the modern sense.
Moreover, the Vindolanda letters suggest that military and commercial activities were not mutually exclusive;
while military shipping is known to have operated out of civilian ports such as Ravenna.
There is no hard evidence for the types of vessel operated by the classis Britannica, although ship types and
crew sizes have been estimated from buildings at Dover (Philp 1981; contra Millet 2007:177) It is worth noting
that there is no explicit reference to the involvement of the classis Britannica in any combative activity,
although their support of campaigning provincial armies or of raiding parties seems likely (see Dio Cassius LXVI
20 1-3).
The occurrence of stamped tiles and inscriptions from Dover, Sussex and Boulogne suggests the organisation
had a cross-Channel communications function with Boulogne being the principle port (Millet 2007 176-7). The
presence of fleet tiles at Beauport Park in Sussex suggests an involvement in iron working and perhaps other
activities.
Milne (2007) has postulated that the Wealden area, bordering a sheltered anchorage fed by the Wealden
rivers and containing a fleet facility in the vicinity of the later shore fort at Lympne, was a centre for iron
production, as well as ship building and fitting out, controlled by the classis Britannica on behalf of the
Procurator Provinciae Britannia in Londinium. This is consistent with its use in the medieval and early modern
periods for ship building, fitting and breaking, carried out at locations such as Smallhythe (Milne 2001; Bellamy
& Milne 2003).
Saxon Shore Forts
During the late Roman period in Britain, a series of military structures, commonly (if erroneously) known as the
Saxon shore forts, were constructed along the south-east coast from Brancaster in Norfolk to Porchester in
Hampshire (Johnson 1976; Johnston 1977; Pearson 2002). Although originally situated adjacent to estuarine
and coastal waters, the changing coast-line has left some forts land-locked, some extensively damaged, and
others have been completely lost to the sea.
Combined with a similar series of structures along the northern coast of Gaul, the shore forts have long been
considered a unified, centralised system built as a one-off response to piracy and coastal raiding, that resulted
from the political, military and economic crises that enveloped the western Roman Empire during the third
century (e.g. Johnson 1976; 1977; Esmonde Cleary 1989:43; Philp 2005: 228).
The interpretation of the forts as purely defensive structures developed from antiquarian scholarship of the
Notitia Dignitatum, a late fourth century Roman text that outlined the litus Saxonicum (Saxon Shore) and its
military defences (Pearson, 2005: 73-74). Much attention in the nineteenth century was focussed on
identifying individual forts in relation to the Notitia (Gardiner 2007: 24; Hingley 2000: 150). However, recent
scholarship in regard to the dating and wording of the Notitia and a re-assessment of the threat of piracy has
questioned the primary role of the forts as a means of defence (Cotterill 1993: 277-234; Pearson 2006; 2005:
77-81)
Recent analysis has reconsidered the function of the forts in the context of the political division of Britain into
a militarized zone in the north and a civilian zone in the south, the economic climate of the third century, and
the re-organisation of the army in the late Roman period (Casey, 1994: 24-25; Pearson, 2005: 76-77; 84;
Hopkins, 1980). Rather than representing a single response to an external threat, recent studies on fort
morphology and building material have shown that the construction of the forts was a two stage process
spanning a period of around seventy years (Allen & Fulford 1999; Pearson 2005: 75-76).
Beginning around c 225, the early forts (Reculver, Brancaster, Bradwell & Walton Castle) were constructed
using more widely sourced materials and to a different plan than the later forts (Allen & Fulford 1999; Allen et
al 2001: 274). Between c 260 and 300 further shore forts (Lympne, Burgh & Dover) joined the network, with
Richborough, Porchester (280s CE) and Pevensey (290s CE) perhaps added as a result of the Carausian revolt
of the late third century (Casey 1994: 24; Allen & Fulford 1999: 181; Pearson 2005: 76; Fulford & Tyers 1995).
The system was further augmented by a series of forts along the Welsh coast (Cardiff, Neath, Loughor &
Caerwent) (Pearson 2005:76; Philp 2005: 227). Together the forts formed a complex military network shipping
supplies from the south of the province, and possibly from Gaul, to the northern military frontier along
Hadrian’s Wall (Cotterill 1993: 236-239; Pearson 2005: 82-84).
Recent excavations within the some of the forts support the notion that they functioned as supply bases. At
Reculver, varied levels of occupation suggest possible temporal changes in function (Philp, 2005: 228-9), whilst
others (Burgh, Brancaster, Porchester & Caister-on-Sea) have revealed evidence for animal carcass processing,
leather-working and industrial activity (Grant 1975: 378-408; Hinchliffe & Sparey Green 1985: 174; Cotterill
1993: 237; Darling & Gurney 1993; Pearson 2005: 83; Moore & Heathcote 2004: 13).
Within some forts (eg Porchester) the presence of open gravelled spaces and the use of ephemeral wooden
internal structures suggests they may have been used to pen animals or store grain; perhaps the annona, the
taxation in kind used to supply the army (Philp 2005; Cotterill 1993: 238; Cunliffe 1975). Similarly, pottery
provides further evidence that the forts operated as bases involved in the transportation of goods. The
evidence shows a distribution of certain pottery types, including types produced in Gaul, along the northern
frontier and along the south and eastern coast of Britain (Allen & Fulford 1999: 177; 1996: 267).
Recent reconsiderations of the function of the shore forts has pointed to a primarily logistical rather than a
defensive role, with the forts operating as fortified ports (Milne 2008). As such, future research involving the
forts needs to further consider the role of the military in facilitating mechanisms of economic exchange
(Pearson 2005: 83). This requires a detailed synthesis on how they functioned as a whole, including the forts
of Wales and Gaul, in the commercial and economic context of the northern provinces (ibid 2005: 74; 85).
This does raise an interesting question; where are the maritime structures that must have accompanied these
fortified ports? Admittedly, many hinterlands have been lost, but what of the land-locked forts? Much focus
on the forts and their interiors, now need to focus on the surviving hinterlands.
It has been suggested that some of the forts were located within or adjacent to areas with low villa density,
usually interpreted as imperial estates (Allen & Fulford 1999: 179). Within this context more detailed research
of the fort hinterlands may help identify road systems linking forts with the interior and the presence of any
vici (small settlements), in order to better understand the relationship between forts and the hinterlands,
particularly with those forts which have suffered extensive damage due to erosion (Good & Plouviez 2007: 10).
The recent identification of the original Roman shore-line and harbour at Dover (Wilmott & Tibber 2009) has
underlined how much the shore-line of Britain has changed between the Roman period and the modern era. A
detailed understanding of how erosion and silting has impacted, and will continue to impact, upon the forts
will help safe-guard them for the future.
Exploitation of the Coastal Fringe
The diversity of the coastal environment provided a range of opportunities for exploitation, some, particularly
salt working sites, can be attested archaeologically while others can be inferred.
Agriculture, fowling and fishing
Large areas of the English coast were salt or estuary marsh and were capable of providing either year-round
rough grazing or summer grazing for cattle, horses and sheep. In some areas this would have added
significantly to the carrying capacity of the countryside and would have provided the opportunity to produce
additional meat and milk products, hides and wool. Archaeological evidence for such activity in the Roman
period is, and is likely to remain, slight though the presence of sea arrowgrass (Triglochin maritimum) at York,
some 30 km from the estuarine Ouse (Dark and Dark 1997: 41), suggests that some indirect evidence might be
forthcoming from archaeobotanical analyses on inland sites.
Vegetable products, such as reeds and rushes, and even barley, which has a higher resistance to salt than
other cereals, could have added to the productivity of the coastal and estuarine marshland, while wildfowling
could have also been a significant additional food source. Evidence for such activity is always likely to be scant
and seems at present to be non-existent, though the examination of fired clay salt working equipment
(briquetage) for the imprints of vegetation may provide some data.
In her review of fish bone evidence from sites around Britain, Locker (2007: 143) notes the absence of fish
traps dated to the Roman period in Britain although examples of wooden traps are known from the Neolithic
and Bronze Age and stone traps and weirs of medieval date have also been recorded. Finds of fish hooks and
other fishing gear in Roman contexts are rare (Locker, ibid.). The fish bone evidence shows a predominance of
freshwater fish over marine species with eels being the dominant fish species identified across the country
(Locker 2007: 154-155). Eels can be taken in freshwater, estuarine and off-shore waters and the extent to
which sea fishing was practised remains unclear. Still more speculative must be the methods for processing
fish, though smoking and salting of saltwater fish for onward trade to inland sites would have taken place near
the coast.
Shell fish, such as oysters which are taken by diving or dredging, are common at inland, or at least riverhead,
urban sites like Colchester, though few were recovered from Heybridge suggesting to the excavators that
oysters were either traded by the coastal population as a “cash crop” or that the trade was controlled from
elsewhere (Atkinson and Preston 1998: 108).
Industrial Activity
Salt extraction sites, known as redhills because of the distinctive colour of the poorly fired briquetage and
former land surfaces which make up the bulk of such sites, are common throughout the coastal marshlands of
Britain. Found principally from the Wash to the Thames estuary, they are also found in Romney Marsh (Kent),
west Sussex/ east Hampshire, Poole harbour (Dorset), the Somerset levels, and Severn estuary. Evidence from
the north-west coast is scant and the general absence of such sites may be the result of coastal erosion or the
competition from inland salt working in Cheshire. Jones and Mattingley (1990: Map 6:4.3) provide a useful
overview of the national picture and their view that the “evidence for the production of salt in Roman Britain
as a whole is impressive, though of uneven quality” still holds true. Since 1990 the number of sites revealing
remains of salt-working and reported in the annual fieldwork surveys of Britannia is roughly two per year.
Various regional surveys of salt working have been published. Some cover all periods of salt working (from the
Bronze Age to medieval), while others are more focussed on a particular period. This reflects the complex
interplay of geographical scale and local topography on the survival of the archaeologically sensitive remains of
a long-lived and probably seasonal industry in a geomorphologically unstable environment.
Essex has been covered by two substantial general surveys (de Brisay & Evans 1975; Fawn et al. 1990), while
the Fens have been the subject of a variety of surveys (Hall & Coles 1994 (multi-period and large-scale); Phillips
1970 (specifically Roman); Fincham 2002 (Roman, eastern Fenland); Potter 1981 (specifically Roman and
covering the central Fens); Gurney 1986 (concentrating on excavated sites in the Norfolk Fenland); Hallam
1960 (Roman sites in south Lincolnshire)). Bradley (1992) describes excavations at a salt working site in
Chichester harbour and places it in its context, while Rippon (2008: 135-137) discusses salt working as an
aspect of trade within the Somerset levels and the Severn estuary.
Throughout the literature, whether general or site specific, a number of key themes emerge: the impact of the
Roman invasion and Romanization on the established salt-working industry, communities and the organisation
of the trade (e.g. Fincham 2002; Gurney 1986), and specifically whether the industry in England was taken over
by the imperial administration or was run by individual entrepreneurs (Salway 1981: 189, 224, 531); the impact
of geomorphological change over time on the industry, and more specifically the reasons for an apparent
decline in coastal salt-working from the 2nd century; the relationship of the coastal salt working sites to the
exploitation of inland salt-working sites at in Cheshire and at Droitwich (Jones & Mattingley 2002: 228); the
relationship of salt working and salt workers to similar industries such as pottery manufacture and metal
working (Salway 1981: 644; Jones & Mattingley 1990: 228).
Mineral and stone resources
Jet, with a geological source off the shore at Whitby, North Yorkshire, and shale from the Kimmeridge area of
Dorset are two examples of the use of mineral products available along the coast. Both were used for jewellery
and, in the latter case, for bowls and even furniture (Allason-Jones 1996, for jet; Calkin 1953 for Kimmeridge
shale). Although jet was worked in York, and not apparently at Whitby, not all artefacts described as jet are
Whitby jet (Allason-Jones & Jones 2001) and the relationship between Whitby and the ‘jet’ artefacts from the
Rhineland, Spain and northern Gaul (Todd 1992) remains little understood, while the trade in shale is perhaps
even less well understood.
Fish
The evidence for fish from Roman Britain comes from 109 sites (Locker 2007), although not all were sieved.
Eel is the most common species, indicating freshwater exploitation, with salmon relatively frequent in the
North and Midlands. Regional trends included the wrasses and sea breams in the south and southwest, which
are also found in the Mediterranean, and Spanish mackerel, imported as salsamenta, although mackerel can
be caught off the south west coast of Britain in warm summers.
The presence of specialist amphorae indicates the importation of garum but there is also evidence for home
production, most compellingly from London (Bateman and Locker 1982). The current picture of fish
consumption in Roman Britain overall is of a continuation in the use of Iron Age freshwater fisheries with an
increase in the use of estuarine and inshore marine species. Locally caught species that were also popular in
the Mediterranean would suggest a ‘Romanisation’ of tastes among the native population, especially from
villas and towns in the south of England, as well as some evidence for imported salted fish or fish sauce.
Religion and Water in Roman Britain and the Northern Provinces
The association between water and the sacred has a long tradition in Britain and north-western Europe(e.g.
Fitzpatrick, 1984; Merrifield, 1962; 1987; Bradley, 1998; Webster, 1997; Parker Pearson, 2003: 179-89;
Osborne, 2004; Hingley, 2006; Kiernan, 2009); particularly the practice of depositing votive offerings such as
weaponry, coins, animal and human bone into wells, rivers, lakes and marshes. Throughout the Roman period
the association between water and belief continued, albeit with some changes in the types of objects
deposited, such as the inclusion of imported samian pottery (Fulford, 2001; Bird, 1992: 86-7; Willis, 1998;
2005), reflecting the availability of new forms of material culture.
Despite this known association between inland water sources and religious belief, less is known in regard to a
direct ritual connection with the sea in Roman Britain, and the evidence, where known, derives predominantly
from military contexts. Altars dedicated to the sea god Neptune have been found along the south coast at the
Saxon Shore fort at Lympne, Kent (RIB 66), on the Antonine Wall at Castlecary (RIB 2149) and along Hadrian’s
Wall (Birdoswald (RIB 1929d; Britannia V 1994, 462-3 no 9); Carvoran (RIB 1788) and Newcastle-upon-Tyne
(RIB 1.1319)). The altar to Neptune from Newcastle decorated with a dolphin, , was accompanied by a further
altar to Oceanus, adorned with an anchor (RIB 1.1320).
Similarly, a portable lead shrine excavated at Wallsend Roman fort (Segedunum), depicted Mercury/Hermes
(the god of travellers) accompanied by a dolphin and a bridled sea-horse, has been interpreted as protection
for those travelling across water (Allason-Jones, 1984: 231). The presence of such altars may be interpreted as
representing the completion of vows following successful military sea crossings from the continent (Hassall,
1978: 41).
Although the large volume of imported goods in the archaeological record of Roman Britain attests to the
importance of maritime trade (Fulford et al, 1997: 124), direct evidence for religious expression by civilian
traders is less secure. An altar stone to Neptune and Minerva from Fishbourne, West Sussex (RIB 92) may be
associated with an inscription of the same date dedicating a temple to the two deities at nearby Chichester
(RIB 92), although the late first century AD date of the two dedications may relate to Chichester’s early role as
a military supply port.
The excavation of altars built into the Roman riverside wall attest to London’s role as an important port
(Britannia 7, 1976, 378, nos 1 and 2), whilst the discovery of a copper as of Domitian (88-9 AD) in the mast step
of the Blackfriars ship attests to a strong degree of superstition amongst sailors (Carlson, 2007). The coin,
showing a figure of Fortuna (the goddess of good luck) holding a ship’s rudder, was placed with reverse side
up, indicating that the goddess was the intended recipient of the offering and its purpose in asking her
protection for the barge (Merrifield, 1987: 54-7).
More formal evidence for religious practice amongst sea-going traders comes from the province of Gallia
Belgica (modern Holland), the location of two Roman period harbours at Colijnsplaat and Domburg (HondiusCrone, 1955; Stuart and Bogaers, 1971; Bogaers, 1967, 1971; Bogaers and Gysseling, 1972). Originally situated
along the estuary of the Scheldt, although now lying off-shore (Birley, 1957), these harbours served ships
trading between the Rhineland, Gallia Belgica, Gaul and Britannia (Halsall, 1978: 42-3). Over 150 altars to
Nehalennia (a guardian or guiding goddess originating from the Rhineland), Jupiter and Neptune have been
found, some depicting Nehalennia with her foot resting on the prow of a ship (Birley, 1957: 173; Bogaers &
Gysseling, 1972; Halsall 1978: 43). Considered to have been dedicated as thanks to the deities after the
successful completion of trading voyages, the altars also highlight the extensive range of goods being traded
between the northern provinces including allec (fish sauce), pottery, figurines, salt and wine (Bogaers, 1971;
Halsall 1978: 44-5).
3. Seafaring
Ships and boats
Five vessels from the Roman period have been discovered around England: the mid-second century AD
Blackfriars I ship (Marsden 1994: 33–91), the late second century New Guy’s House boat (Marsden 1994: 97–
104), the third century AD Barland’s Farm boat (Nayling et al., 1994; McGrail & Roberts 1999; Nayling &
McGrail 2004), the late third century AD County Hall ship (Riley & Gomme 1912; Marsden 1994, 109–28), and
the St Peter Port ship of similar date (Rule & Monaghan 1993). Three of the vessels were found in London
(Blackfriars, New Guy’s House and County Hall) in riverine contexts, as was the Barland’s Farm boat. Only the
St Peter Port vessel from Guernsey was found in a maritime context. Four were believed by the excavators to
be sea-going; Blackfriars, County Hall, Barland’s Farm and St Peter Port.
Blackfriars, Barland’s Farm, New Guy’s House and St Peter Port represent a robust carvel-style indigenous
vessel-building programme, the so-called ‘Romano-celtic’ tradition, all of which belong to McGrail’s (1995:
143) sub-group B, which he regarded as sea-going. This tradition existed alongside the native logboat and/or
hide boat traditions, both of which predate the arrival of the Romans (Caesar, de Bello Gallico, iii, 13; de Bello
Civile, i, 54), but continued in use throughout the period. The Hardham II logboat for example dates from the
later Roman period.
These two very different techniques would clearly have utilised very different methods and construction sites
(Milne 2008). In addition, the County Hall ship provides evidence for the introduction into northern Europe of
Mediterranean ship-building techniques, although it was built from trees grown in south-east England
(Marsden 1994: 124). While the function of this particular vessel remains unclear (Marsden 1994: 125–7), the
introduction of Mediterranean methods is likely to have been associated with warships.
Besides hide boats and Romano-Celtic ships, Caesar (de Bello Gallico, iii, 9, 14; iv, 22, 25) also describes the use
of Mediterranean-style warships in British waters in the first century BC, some built on the Loire and others
possibly brought from the Mediterranean. However, the earliest vessel of any tradition discovered from
Roman Britain dates from the mid-second century AD, representing a lacuna of two centuries for which we
have no archaeological evidence. Furthermore, owing to the nature of the construction material, other than
the Broighter gold model from Ireland, there is virtually no archaeological evidence for hide boats. There is
documentary and iconographic evidence, some of which dates from later periods, thus indicating a continuing
tradition.
The evidence for Britain from literary sources and maritime-related pictorial representations is minimal
(Ellmers 1978). A coin of Cunobelin, the king of the Trinovantes, found at Canterbury depicts a ship
(Muckleroy et al 1978), whilst the Arras medallion (Askew 1980: 54), an intaglio recovered from the Thames
(Henig & Ross 1998), a silver denarius of Carausius in the British Museum, and a billon quinarius of Allectus, all
depict Roman warships (http://www.kenelks.co.uk/coins/carausius/carausius.htm). In addition, various tiles
stamped CLBR record the location of bases of the Classis Britannica and provide some indication of its
activities (Cleere 1977), whilst a floor tile in the British Museum, probably from London, bears a graffito
depicting a lighthouse.
Apart from Caesar who, in addition to his accounts of ship types, describes his invasion passages in great
detail, documentary sources are tantalizingly brief and short on detail. Pliny the Elder (Nat. Hist., iv, 104)
refers to hide boats framed with withies in connection with the trade in tin. Strabo (Geogr., iv, 5, 2) tells us
that passages were habitually made to Britain from the Rhine, the Seine, the Loire and the Garonne, adding
that grain, cattle, gold, silver, and iron, together with hides, slaves and dogs were exported from the island.
Tacitus (Agr., 10, 28) records the mutiny in AD83 of a cohort of Usipi, who sailed three liburnian galleys round
the north of Scotland, ultimately losing them through their nautical incompetence and briefly notes the
circumnavigation of Britain by a Roman fleet the following year.
Vegetius (de re Mil., iv, 37) offers an obscure reference to forty-oared scaphae exploratoriae (scouting skiffs),
which the Britons call picati. Dio Cassius (Roman History, lx, 19–22) in his account of the invasion of AD43
gives virtually no insight into the maritime context of the campaign. Similarly, the sources for the recovery of
Britain by Constantius Chlorus after the Carausian secession are similarly uninformative, offering only the
unenlightening information that both invasion fleets sailed close-hauled in bad weather conditions (Aurelius
Victor de Caesaribus, 39; Eutropius Breviarium, ix, 21–2; XII Panegyrici Latini, x(ii), 11–12 and viii(v), 6–7; 12–
20).
Rig and performance
That Romano-Celtic ships were rigged for sail is evidenced by their forward-located mast step. There is no
direct archaeological evidence for the sail type or sail material, apart from a second/third century AD
monument in the Rheinisches Landesmuseum, Trier, and a third century floor mosaic from Bad Kreuznach,
Rheinland-Pflaz, which show a battened square sail, and Caesar’s report that the sails of the Venetic craft were
made of leather. The possibility of square sails, spritsails and lugsails has been explored in various studies
(Marsden 1994: 70–4; McGrail & Roberts 1999: 138–9; Grainge 2002: 32–5; 43–4). Studies of sailing
performance, stability and load-carrying capacity have been published (Marsden 1994: 193–9; McGrail &
Roberts 1999: 141–2; Grainge 2002: 113–20; 125–6); inevitably in the absence of firm evidence for the type of
rig there is a degree of uncertainty about the results.
Documentary research and the sea trials of the reconstructed Athenian trireme have provided more
information about the rig, rowing configuration and performance of Mediterranean-style warships (Morrison
& Coates 1996; Morrison et al 2000). However, extrapolating performance data from the Mediterranean to
Mediterranean craft in British waters is not without its uncertainties.
The archaeological evidence pertaining to the rig of hide boats is virtually non-existent as is the archaeological
evidence for their performance. While it is very likely indeed that the larger sea-going versions were rigged for
sail, only the Broighter model indicates the possibility of a square sail. Thus no contemporary data exists
which might allow even tentative conclusions as to the performance of Roman-period hide boats.
Seamanship, pilotage and navigation
The navigation of the period was essentially non-instrumental (McGrail 1987: 276–84). McGrail’s (1983) study
probably continues as the definitive study of cross-Channel seamanship in the late Iron Age and may validly be
extrapolated into the Roman period. Passage-making would have been in great measure influenced not only
by the likely need for a favourable wind, but also by twice daily ebb and flow of tidal streams, making it
desirable where possible to work the tides.
Seafarers would have used the direction of the wind, the swell pattern, land marks, both man-made and
natural, the sun and stars to find their way and would have been able to read the surface of the sea to detect
the presence of underwater shoals and safe channels, as experienced modern yachtsmen continue to do
(Sleightholme 1970: 107–13). Among man-made marks were the lighthouses which marked the entrances to
Dover and Boulogne (Suetonius, Calig., 46), one of which (at Dover) is extant, the other (at Boulogne) is
recorded as surviving until 1644 (Diderot & D’Alembert 1751-1772: 489); each were probably one of a pair
sited on headlands either side of each harbour entrance.
4. Maritime networks
- International connections, trade, exchange, migration, naval conflict, Empire and colonisation etc
Roman Military/Roman Navy- A very important topic in its own right
The role of the Classis Britannica (Mason 2003) is of particular importance in the first and second centuries.
Initially it would have organised most of the coastal and cross-Channel traffic. This would have changed in the
later Roman period with the development of the so-called Saxon Shore forts, which are probably better
described as ‘fortified ports’. (Milne 2008)
“Trade” versus “Military Supply”
There are lots of snippets about “trade” and “military contracts” in pottery reports, particularly synthetic
reports about towns or regions, plus conferences, festschrifts, BAR’s and the Journal of Roman Pottery Studies
from the 1970’s onwards. A fair amount of re-inventing the wheel is possible amongst the new breed of
postgraduates who are not always aware of the earlier works.
Merchandise from many parts of the Empire seems to have reached Britain via a complex system of transshipment centres, using river barges and sea-going vessels, up the Rhone and the Rhine, then across the
Channel to the Thames. The overwhelming concentration of inscriptions related to the shipment of goods on
the Rhône-Saône axis highlights the dominance of this route as the principal commercial axis of Gaul
(Middleton 1979: 82; fig 1). Evidence from 124 altars, found at the trans-shipment centres of Colijnsplaat and
Domburg on the River Scheldt, confirm contact with British traders. These centres served ships trading
between the coastal regions of Gaul and the east coast ports of Britain on one hand, and the Rhineland and
Gallia Belgica on the other (Hassall, 1978, 43-4; Middleton 1979: 95; Milne 1990).
The complete absence in northern Europe of ships built in the Mediterranean, and the preponderance of
native craft probably reflects the largely terrestrial and riverine contexts in which the majority of these vessels
have been discovered. Although the presence of Mediterranean ships cannot be discounted, the
predominance of local ships and boats seems a good indication of the types of vessel that frequented the
major ports of northern Europe. This appears to confirm that long-distance trade was conducted via the
inland waterways of Gaul (Strabo IV 1.2; 1.14), which were navigable along all the main axes of communication
(Middleton 1979: 82), rather than open-sea voyaging around the Atlantic coast. Avoidance of long-distance
sea voyaging is further supported by literary evidence, which reinforces the notion that sea voyaging, even in
the relative safety of mare nostrum, was a dangerous undertaking that should be avoided (see Suetonius,
Gaius 47; Claudius 17; Dio LIX 25.2).
One of the striking features of the province of Britain is the gradual decline of imports relative to home
produced goods, in the archaeological record, from a peak in the pre-Flavian period which Fulford (1978: 62)
believes reflects the developing Romanization of Britain rather than a progressive decline in long-distance
trade. This decline might suggest that by the third century the material culture of Britain had equalized with
that of the Empire, so Britain was no longer dependent on imports (Fulford 1984: 137; Millett 1990: 162).
In addition to these known vessels, contemporaneous amphorae recovered from the Little Russel Channel,
Guernsey, and deposits from Richborough (Lyne 1999), Nournour on the Isles of Scilly (Fulford 1989), Herd
Sand at South Shields (Bidwell 2001), and from Hartlepool Bay (Swain 1986) have been interpreted as remains
of either ships or cargoes (see Walsh 2002). The famous Pudding Pan site is now thought to comprise at least
two wrecks. The main assemblage dating from c 175-195 AD comprises central Gaulish samian, a range of
amphorae, and a solitary ARS bowl, while the later first century assemblage (c 65-85 AD) comprises a number
of mortaria and amphorae recovered between the Oaze Deep and Pan Sand. There is also a small assemblage
of unprovenanced early third century material from the same locality (Walsh 2006; see also Hartley 1977: 6).
An assortment of maritime finds from around the British Isles provide further tantalizing hints of other possible
sites, although at least some of these are likely to represent casual losses either thrown or lost overboard
rather than a shipwreck or cohesive archaeological site. However, concentrations of material discovered in
similar locations over time do warrant closer inspection. For example, there is anecdotal evidence that iron
anchors and planking recovered in the nineteenth century from the West Caistor marshes near Caister-byNorwich came from a Roman boat (Fryer 1973: 269). For centuries, a whole variety of Roman artefacts have
been, and continue to be, recovered from maritime contexts in northern European coastal waters including
amphorae (Galliou 1982; Sealey and Tyers 1989; Harmand 1966; McDonald 1977: 24; see Fig. 8), pottery
(Monaghan 1989; 1991; Pownall 1778), coins (Dean 1984: 79), ingots (Craddock & Hook 1987; L’Hour 1987),
anchors (Boon 1977a & b; Cook 1971; Dean 1984: 79; Marsden 1990: 71; Markey 1991; 1997), military
equipment (Bidwell 2001), roof tiles (Spurrell 1885: 281-4), and brickwork (Pownall 1778: 282). However,
although individual finds have been researched and occasionally published there has been no synthesis similar
to the corpus of artefacts found off the French coast (see Galliou 1982).
Road network – connections with navigable rivers – Lincoln, York, Catterick, Hadrian’s Wall,
Roads/facilities at tidal heads – ships in, barges out
Pottery evidence from Exeter and Plymouth
5. Maritime identities and space
The landscape of Roman Britain was not a passive backdrop, it was a place that was moulded by human action
and at the same time influenced further action and interaction. The on-going and reflexive relationship
between people and surroundings provides traces within the archaeological record which, in turn, give
opportunities for contextual interpretation of identities and the use of space.
To date much of the study of Roman Britain has focused on the process of Romanisation (Revell 2009) and
functional factors such as the use of resources and developing environmental and economic aspects (Petts
1998). Roman archaeology has, perhaps, been slower to adopt more interpretative approaches that consider
agency and meaning. This is not to overlook more traditional approaches to ‘space’, such as the control of
territories and associated landing places and coastal resources, which would have a bearing on identity, but to
advocate a broader interpretation of evidence drawing on recent work and approaches.
There is a need to look at the Roman seascape and coastscape as an active space as it was experienced. These
areas or environments did not have a single ‘meaning’ as this is highly individualistic. However, it is possible to
explore the potentially multi-level and reflexive nature of space linked to the marine and maritime zone
through the evidence base from the Roman period (see Ingold 1993).
When seeking to understand identity in the Roman period there are challenges related to the social, political,
economic and cultural ‘space’ of the province which were based on unequal power relationships which
affected all aspects of life (Revell 2009). In terms of maritime and coastal environments this raises a wide
variety of research questions related to the significance and/or ‘difference’ of these environments and how
this may have affected the varying experiences of those living within, or passing through, them and the
negotiation of social identities. What effect did the Roman invasion have on social identity both for the
indigenous population and for the invader? How did this alter the individual’s perception of the
sea/coastscape in which they lived?
Case Study: The Solent in the Roman period – Magnus Portus
As in the present day the Solent region in the Roman period had a very maritime focus. Tomalin (2006: 49)
explores maritime aspects of villas in relation to safe harbours and anchorages and advocates that the
maritime interests and focus of the Roman populations should not be under-estimated.
The intertidal environment includes the important industries of salt-working and oyster fishing, as well as
other uses including seaweed gathering, sheep grazing on marshes, and fowling. Local maritime activities
included fishing and ferries engaged in the movement of people and resources, such as stone, from the Isle of
Wight to Fishbourne villa. In addition to local and regional maritime traffic and communications, international
vessels would have anchored in areas such as the Mother Bank and Yarmouth Roads or proceeded to port
facilities at Clausentum (Beattie-Edwards 1999).
Transport infrastructure linked roads to the tidal river systems and harbours. In areas where shallow water
prevented larger boats from reaching sites directly, such as at Fishbourne villa (MoLAS 2007:22), goods were
transhipped onto smaller barges (Rippon 2008). The crews of ships anchored in the Solent for extended
periods (during transhipment) would have required provisions taken to the ships.
Intensity of use of the sea and rivers enhanced their strategic importance for trade and transport, which in
turn shaped the maritime environment not only physically, in terms of presence of shipping and facilities, but
also perceptually; how it was experienced by those within it. From the land the presence of sailing and
anchored ships (or even wrecked or beached ships) would have a significant impact within the visually
contained waters of the Solent. Equally important is the view from the sea; how did sailors find their way
around? How well did they understand the maritime world and thereby contribute to the construction of their
identity(ies)? How did this vary between ‘native’ and Roman? Between civilian and military? How did these
perceptions vary over the lengthy period of Roman occupation?
High volumes of international and regional shipping would have brought with it peoples, goods and ideas from
across the Empire. Inevitably this would have transformed the ‘outlook’ of those in the coastal areas and
beyond. A diverse range of sources and material culture inform us regarding relationships with and
perceptions of water and the maritime environment. The relationship of major temples with the maritime
environment, such as the temple on Hayling Island which lies over the remains of its Iron Age predecessor
(King & Soffe 1994), augments the location’s significance in terms of activities in the area and access to the
temple for those arriving by sea.
The influence of the sea and maritime themes is reinforced in mosaics (eg Fishbourne and Brading) and on
coins (Orna-Ornstein 1995). Neptune (god of water and the sea) and other water deities were also influential;
a find from the River Hamble suggests another marine relationship. A lead curse tablet (defixio) calls for
Neptune and Niskus (likely to be another water deity) to take revenge on those that had stolen coins from
their owner (Tomlin 1997). The practice of ritual deposition in water continues a long heritage; water has
always been seen as a liminal zone, from Bronze Age Scandinavian communities through to modern folklore.
Shipboard space and identity
Besides coastal communities, those travelling onboard vessels experience maritime space, living within a
mariner-based community. Some journeys may have involved a short voyage, such as across the Solent, while
others would have travelled considerable distances. The experience of the mariner or passenger in terms of
the agency of space is an area of research that it not particularly advanced. More work has been done on the
required skills and practicality of sailing earlier vessels (McGrail 1987), and on the experience of maritime
space in the broader sense of requirements for understanding the sea, weather and landing places.
The extent to which the division of space, roles and materials onboard vessels effect and affect social actions
and identity is not well developed. This is not surprising owing to a greater focus on nautical technology and a
lack of vessel remains from Britain. However, there are opportunities to review the division of space, goods
and activities that shaped identities based on reconstructed examples such as the Blackfriars (Marsden 1994)
and Guernsey (Rule and Monaghan 1993) vessel remains.
Social space on vessels was dictated by the functional space related to cargo and vessel performance. The
relations between crew members may be reflected within material culture and non-functional activities. The
importance of religious beliefs to those onboard ships that traversed the often dangerous and unpredictable
seas is demonstrated by the presence of shrines, often sited on the ship’s quarter deck, an area which
continues to have special ceremonial significance (Reilly 1975). This is supported by the discovery of the
numerous altars at Domsburg and Colijnsplaat dedicated by seamen to a safe passage (ref)
Research Agenda
1. Coastal Evolution
The prime concern for maritime landscape archaeologists is to establish the form and navigability of coasts and
estuaries in the Roman period (see for example Grainge 2006; Perkins 2006). A primary objective is therefore
the compilation of a working curve for historic sea-level change relative to the land, so that the high- and lowtide values can be ascribed to the coastal and estuarine contours, ideally on a century-by-century basis. A
closely associated changing coastal morphology model should be developed in parallel to this long-term
research aim: sand banks form, promontories erode, estuaries and lagoons silt up, the tidal head of rivers alter,
the extent of useable open foreshore changes. The Romney Marsh Trust has been facilitating just such
research in their study area for over twenty years, in an exemplary fashion (e.g. Eddison 1995; Eddison et al
1998; Long et al 2002). Such models will help predict where deep-water ports, fishing settlements, shipyards
might or might not be located, and thus where vessel fragments and hulks might also be found.
A major mapping project is therefore proposed, plotting the coastal changes at 200-year intervals throughout
the Roman era. The series of coastal models would be compiled from regressive map and chart analysis (for
land and submarine contours), soil survey maps, sedimentary analysis, modified with borehole logs, dated
excavation data and inter-tidal zone survey (eg Allen et al 2004). This approach could be piloted in study areas
such as Romney Marsh or the Wantsum Channel, after which the coastal map series produced would serve as
the chronological base maps for the regional HER centres, so that the subsequent plotting of finds and sites
would be directly related to the ‘correct’ contemporary coastal morphology. Once the pilot studies have been
produced the revised methodology could be extended to other areas of the coast. It is accepted that such
studies will not be without their problems and uncertainties, owing to the many variables that need to be
taken into consideration, but it is only by identifying and addressing the problems that progress on the related
research themes will be made (Milne 2008).
Sea-level change and the shift in the coastline are not the only geomorphological events to occur in a coastal
context. It is clear that sand dunes systems are complex entities prone to instability and sudden, large-scale
shifts. This may have important consequences for recognising, dating and conserving sites in these areas.
Further research on the geomorphology of these sand dunes and their movements should be a priority (Petts
& Gerrard 2006).
Research Framework Priority: Collation of data permitting reconstruction of Roman period
palaeogeography and identification of areas where research is needed.
The DEFRA-funded shoreline management plans (SMPs) consider options ranging from non-intervention or
maintenance, to managed retreat or advancement of the shoreline, all of which have implications for Roman coastal
archaeological sites. In some cases predictions of uncontrolled coastal erosion is highly perturbing and must
determine our priorities. Hard coastlines with cliffs may be relatively easy to assess archaeologically but there is
surely much attention yet to be given to rias and estuaries where post-Roman sea-level rise has the propensity to
engulf, conceal and preserve. Reclaimed coastal inlets, channels and wetland are a further resource for Roman
maritime studies. The case has been well demonstrated on Somerset’s ‘North Marsh’ and in Romney Marsh; the
Wantsum Channel, Brading Haven and the East Anglian coast clearly demand closer attention.
So far, the coastal process units have been examined in terms of shoreline dynamics and coastal engineering but few
have received attention from an archaeological view. The coastal archaeological audits commissioned by English
Heritage are an excellent start and they offer a means of focussing attention on coastal archaeological remains above
mean low water mark. We must also address the effects and the missed opportunities posed by dredging in
Britannia’s long sinuous riverine highways such as the Thames, the Trent and the Tyne.
Controlled archaeological trawls in the offshore zone have identified significant Roman anchorage debris on the floor
of the Magnus Portus zone of the Solent. A recent review of coastal and maritime villas identified some forty historic
natural havens and offshore anchorages between Margate and Plymouth Sound; a surprising number are attended
by coastal villas such as Folkestone, Eastbourne, Littlehampton, Pagham, Langstone, Brading, Newport, Gurnard,
Bowlease, Weymouth, and Honeyditches (Tomalin 2006). Seabed assemblages such as these have a special
significance as ceramics discarded from anchored ships frequently include tell-tale and exotic items that can vanish
from the archaeological record once they have passed across the gang-plank.
2. Maritime settlement and exploitation
Such regional models will also inform coastal settlement studies: some coastal areas have changed so
dramatically, that this will be reflected in the location and form of associated towns. The varying plights of the
so-called Saxon Shore forts amply illustrate the dramatic extent of maritime settlement dynamics. Added to
this a change in sea-level relative to the land, in the form of a marine transgression, would have impacted
upon the viability of the associated harbours so crucial to the maintenance of Britannia's link with the
European mainland and to Rome itself. The relationship of each of those ports to the changing coastline and
tidal levels needs to be explored and established. The time is now ripe for beginning a new, wider study,
looking at all the towns as a group, set in their full coastal, economic, social, political and topographic context.
There are other issues that the study of such maritime centres also needs to consider. An obvious research
priority would be the form and development of harbour works: with the exception of Dover, few ports can at
present contribute to such a debate on archaeological grounds alone. The situation in London and elsewhere
on the Thames was much the same some 30 years ago: it was only the extreme pressures of urban
redevelopments there that exposed so many ancient harbour sites over such a relatively short time period
(Milne 1985; 2003). Should the demands for flood defence or urban renewal increase elsewhere, then a
similar programme of developer-led waterfront excavations might ensue: until then, there is considerable
scope for exploratory survey and evaluation work to establish the possibilities.
Other research themes requiring attention include consideration of maritime urban topography (Bill and
Clausen 1999), maritime buildings, which can include churches, as a recent paper has shown (Cohen 2008),
lighthouses (the Dover pharos is a prime example) and other seamarks (Naish 1985). A revealing study of
discarded ballast reused as building material was recently conducted in King's Lynn, in Norfolk (Hoare et al
2002): similar exercises might well be undertaken for other periods and in other regions with profit.
Changing maritime industries
Once the coastal models have been developed and broadly established, the navigability of the associated
estuaries, creeks, rivers, lagoons and open coast can be determined. As discussed, this will also facilitate the
discovery of the boat and ship-yards that must have been a key feature of this coast. For example,
geoarchaeological research has demonstrated that there was once a large sheltered lagoon on the Kent/Sussex
border, the mouth of which was once protected by a fort of the Classis Britannica, at Lympne. Given that
archaeologists have established many iron-working sites in the timber-rich Weald to the north, as well as the
presence of slag-roads leading towards the now silted-up lagoon, it is not an unreasonable suggestion that the
fleet vessels (requiring prodigious amounts of timber and ironwork) were built there (Milne 2000).
The location of vessel-building sites changed in response not only to sea-level rise or harbours silting, but also
to major changes in shipbuilding technology like the transition from small ‘native’ craft to Roman carvel
construction. Surprisingly few shipyard sites have been found. In addition to shipyards, with their associated
smithies, rope-walks and sail-makers, there are other maritime industries for which archaeological research
can make a real contribution. The role of fishermen and the fishing industry has often been overlooked in the
development of the economy and of the region's first towns. The archaeological study could examine
settlement sites, artefacts such as fishing boats, fish hooks and net weights (Steane and Foreman 1991), and
the biological remains of fish, shell fish and other marine creatures (eg Riddler 1998). Sea-salt-making is
another coastal or estuarine industry of some importance (the fish had to be preserved), and there is evidence
ranging from the Iron Age through to the medieval period (Topping and Swan 1995; Champion, 2007, 110,
Ridgeway 2000), while the Kentish copperas industry also merits attention here (Allen 2002), as does the
Roman pottery industry (e.g. Monaghan 1987)(Milne 2008).
Changing maritime defences
Defence in this context has two meanings - defence against the sea, and defence against sea-borne invaders.
The history of the defence against natural elements is directly related to studies of coastal morphology and
sea-level change, with the dramatic addition of major storm events. Once again, the teams working in the
Romney Marsh area are the leaders in the search for how much land has been won from or lost to the sea
(Eddison 2000). Our understanding of the landscapes of defence must again be considered with due regard to
changing coastal morphology.
Marine mammals & fish
Pitifully few Romano-British sites have produced evidence of fish consumption; in order to establish more
realistic trends in fish consumption for Roman Britain this area needs further exploration (Locker 2007). This
can be achieved through analysis of data from past excavations and through recommendations for future
work. The data from all previous Roman levels/excavations could be examined to ascertain what, if any,
sampling for fish bones was conducted.
Could the perceived absence of fish remains be attributable to a) no sampling, b) inadequate sampling using
too great a mesh size, c) random sampling, or d) sampling carried out to appropriate standards but no fish
remains were found. On fish-free sites, were other small bones such as those from small mammals, birds or
rodents recovered? If so, these data should be compared with data from excavations where fish have been
found. Future work would include specific sampling strategies to target the recovery of fish remains, such as
bulk samples sieved to 1mm with a subsample checked to 0.5mm. Advance direct liaison with commercial
archaeologists would be essential to ensure implementation of sampling strategies.
Harbours
Some substantial structures, such as wooden landing stages, have a limited life and Roman period remains are
unlikely to be found, other than during excavation at known locations such as Caerleon and Chester. Other
methods of unloading, such as beaching, leave no trace in the archaeological record. The problem is well
demonstrated by the following for which no evidence survives: the quays at Bath, probably equipped with
cranes for loading large blocks of Bath stone; a landing place serving the important villa at Keynsham; another
for the walled settlement at Gatcombe, and a significant port at Sea Mills. In addition, the recent discovery of a
large villa complex near to Bradford on Avon (Corney 2002) and the presence of a medieval quay near to the
town bridge suggest probable navigability of the river to that point, yet no Roman remains have been found.
In these cases, the remains of Roman riverside structures are unlikely to have survived: the development of
later quays at Bath; the construction of wharves for Fry’s chocolate factory at Keynsham, and the development
of a ‘wet dock’ at Sea Mills. As for Gatcombe, it lost its significance at the end of the Roman period, and any
landing place has long since been eroded by tidal action. The only surviving remains of any Roman maritime
structure is a small piece of walling at Sea Mills that may, or may not, have formed part of the port wall.
(Jones 2009: 48).
The flat-bottomed style of the St Peter Port, Blackfriars and Barland’s Farm vessels suggests that they did not
need to berth in formal harbours and were possibly designed for regions where the tidal range is so great that
harbour structures are only useful at certain states of the tide, or indeed regions without harbours. So, we
know there was a Roman harbour at St Peter Port, but it need not have been more than a beach protected by
reefs. This, coupled with the loss of the Roman coastline, suggests that the search for more formal harbour
structures may be a fruitless one.
Medieval records of royal voyages to the continent amply illustrate that significant ‘hards’ and landing-places do not
necessarily demand attendant wharves, ports or roads. Surveys of the Thames foreshore, the lower Itchen, and the
Solent and Severn estuary coasts all demonstrate the remarkable quality of the Romano-British archaeological
resource concealed within the inter-tidal zone. The Wootton Quarr survey has demonstrated that the practice of
beaching and unloading Roman craft without the facilities offered by a pier, jetty or wharf, can carry the propensity
of depositing and preserving a significant array of lost cargo and goods in the inter-tidal silts. It is a revelation that
now concerns all manner of simple sheltered landing places on Britannia’s coastline. While Ptolemy’s Magnus Portus
is just beginning to reveal its seabed secrets, in terms of artefact scatters, we have yet to address his other named
anchorage off the Yorkshire coast at Safe Haven Bay.
Classis Britannica
Address the lack of archaeological evidence for its activities, even in what is assumed to be its heyday in the
second century AD. What was its role? Sphere and range of operation? Over what time period? Evidence for
military activities on civilian sites?
Military anchorages and landing sites.
The submerged dimension of military fortresses and shore forts, such as Deva, Glevum, Isca Dumnoniorum,
Camulodunum and Durovernum, deserve particular attention. At Brancaster the estuarine channel was perceptibly
migrating during the twentieth century when Steers (1964: 358-67) observed the presence of preservative wetland
environments nurtured by the behaviour of protective dunes and shingle spits. Outstanding finds, like the bronze
gladiatorial mask from the bed of Tyne, speak eloquently of un-investigated sub-tidal contexts, while the daunting
assemblage of skeletal remains associated with the riverine context of the Battersea shield is a further example of
under-investigation.
Portchester’s great natural anchorage in Portsmouth Harbour, vulnerable to modern naval and commercial
developmental pressures, certainly calls for targeted archaeological attention. The bed and margins of the Elai at
Cardiff, below both the shore fort and the Ely villa, call for similar attention, while the proximity of the Wantsum
channel to Richborough and Reculver elevates its archaeological potential (see Hardman & Stebbings, 1940-2)..
Although SMPs have recognised the need for some archaeological interventions, widespread weaknesses and
inconsistencies have been identified (Firth 1998).
Riverine highways
A review of the archaeological content of all river management plans is certainly required. The presence of
Neidermendig lava querns on Thames sites like Chertsey and Staines hint at Roman riverine navigation yet the
depositional archive contained within the riverbed remains under-investigated and is ever vulnerable to destruction
by unmonitored navigational dredging. Riverbed evaluations adjacent to significant Roman riverside settlements
such as Staines and Dorchester (Oxon) must also pay dividends.
The Trent must also have been an outstanding Roman riverine highway from the Humber estuary deep into the heart
of Coriotauvian territory accommodating riverside settlements at Segelocum, Ad Pontem and Margidunum. Historic
navigation problems probably required Romano-British craft to lay-off in their approach to the Trent mouth and the
Ouse route to York so the potential survival of anchorage strews at and west of Horkstow should be investigated.
Other riverine highways that warrant evaluation include Belisama where Ptolemy appears to identify the estuarine
approach via the Ribble towards Ribchester, and the silted and drained course of the Parrett on the Somerset coast,
due to its relationship with a number of adjacent villas including Huish Episcopi, Low and High Ham, Bawdripp and
Puriton (Tomalin 2006). The relationship between the Parrett and the chain of Silurian coastal landing places and
villas has been noted (Tomalin, ibid), yet there remains a significant lacuna of information concerning the mouth of
the Axe at Uphill, overlooked by the Brean Down coastal temple, where the putative Roman road from Charterhouse
meets the Mendip seaboard. On the Silurian coast estuarine approaches via the Afon Tywi to Moridunum
(Carmarthen), the Llwchwr to Levcarum, and the Neath mouth to Nidum all call for sub-tidal and inter-tidal
evaluation where they offer important landing and lay-over points on a coastal navigation route reaching to the Irish
Sea.
On the Channel coast, the implicit relationship between the riverine highway of the Arun and the Wigginholt iron
industry requires investigation. Early records of log-boat discoveries in the floodplain hint at particular
archaeological potential here.
Ships and cargoes offshore
It is surely significant that, with the exception of some abortive sorties over the Pudding Pan wreck site, little or no
prospection has actively sought the presence of Romano-British craft. Where this was specifically pursued at
Yarmouth Roads in the Western Solent, the result was the prompt identification of a significant anchorage strew
(Tomalin 1997).
In addition to those sites already named, there remain significant offshore navigational hazards upon which ships
throughout history have met their fate especially the Goodwin Sands, but also Bognor Rocks, Shelly Rocks, Kingmore
Rocks and Winters Knoll Shoal that impede safe approach to the mouth of the Arun. Third century coins recovered
off the Needles at the Western approach to the Solent point to further Roman ship losses.
Rias, inlets and civilian harbours and anchorages
Although Ptolemy’s map of Occeanus Britannicus, shows only three river inlets west of the Magnus Portus,
seemingly the Exe, Plymouth Sound and Carrick Roads at Falmouth, the south-west peninsular coastline
abounds with sheltered rias and landing beaches that provide many of Britannia’s finest natural harbours
although there is a dearth of known conventional Roman settlement. There is a danger that the inter-tidal and
sub-tidal archaeological potential of these locations will elude the archaeological agenda for Roman Britain
and this needs to be addressed. The Radipole inlet on the Durotrigian coast is a good example of high
development pressure, a destructive dredging threat and little if any archaeological mitigation.
The off-shore location of the two shrines at Colijnsplaat and Domburg, coupled with the changes to the Roman
coastline might explain why few securely attested coastal shrines and temples are known from Roman Britain
(Fulford et al, 1997: 122-7). Although some Roman period coastal religious structures do survive, for example
the fourth century temple at Brean Down, Somerset, located on a headland projecting into the Bristol Channel
(ApSimon, 1965: 199) and Nor’ Nor, Isles of Scilly, they are little understood and, in the case of Nor’ Nor, its
interpretation as a religious shrine has recently been questioned (Fulford, 1989). Future research needs to
focus on better understanding these surviving structures in relation to the sea. This, allied to surveys to
identify the original Roman coastline, may help to identify where probable shrines or temples may have stood.
In light of known patterns of structured votive deposition relating to inland religious water sites, a reexamination of coastal material culture finds may also help to identify Roman period religious and ritual
behaviour in relation to the sea.
3. Seafaring
Ships and boats
The temporal distribution of these largely serendipitous discoveries is erratic. There are periods of extensive maritime
activity around Britain for which there is a considerable hiatus in evidence for these most complex and most obvious,
largest maritime artefacts (Arnold 1978: 32). This lacuna in maritime evidence from around the British Isles spans several
hundred years, from the pre-historic Humber boats to the mid-second century AD Blackfriars I ship (Walsh 1998: 25; Adams
2001: 307). Consequently, we know more about the minutiae of the so-called ‘Romano-celtic’ or ‘Gallo-Roman’ boatbuilding traditions (see Ellmers 1969; Marsden 1967; 1977; Arnold 1978; de Weerd 1978; 1988) than we do about the
transition in maritime transport from the Bronze Age through the Iron Age to the Roman era (Adams 2001: 307; see also
Johnson 1999: 21).
While finds of ships of the period will continue to be fortuitous until such time as pro-active research is
undertaken, significant research objectives would be met with the discovery of:
more ships built in the Mediterranean tradition
any archaeological vestiges of hide boats
finds of Romano-Celtic craft from early in the period – or even from the late pre-Roman Iron Age, thus
confirming the documentary evidence of Caesar’s description of the Venetic ships as an indigenous
non-Roman ship-building tradition
any finds which parallel contemporary ship-building traditions from the Rhine:
inland Romano-Celtic craft of McGrail’s sub-group A (McGrail 1995: 143)
warships of the Mainz-type, which appear to have adopted indigenous, rather than
Mediterranean, construction methods (Höckmann 1986: 392–5).
Rig and performance
Studies of the rig, oar configuration and performance of ancient (Mediterranean-style) warships are
reasonably secure in so far as they have been validated by the sea trials of the reconstructed Athenian
warship. With caution they may be extrapolated to Roman warships operating in British waters, but could
benefit from further archaeological data from British waters.
By contrast, studies of the performance of Romano-Celtic ships have thus far been theoretical only. We may
accept the conclusions as to stability and carrying capacity with some confidence; studies of performance
under sail, in particular of windward ability, depend on large assumptions as to the size and type of the sail and
as to the hull’s leeway resistance.
The following would improve our understanding of the issues involved in studies of sailing performance:
further archaeological evidence of the sail type and size (unlikely, since the rig is almost always
entirely lost)
wind tunnel data on the various probable sail types
tank-testing of hypothetical models with different configurations of rig and sail
The possibility of building a full-scale replica hull, capable of being rigged with different types and sizes of sail,
should not be discounted.
As for hide boats, without further archaeological data, it is not realistically possible to go beyond existing
conclusions about their performance (McGrail 1987: 184; Severin 1978: 288–91).
Seamanship, pilotage and navigation
It is reasonable to assume that the offshore tidal pattern of the Roman period was much the same as that
which applies today (Devoy 1982: 85). However, this cannot apply to inshore waterways or to navigable rivers,
since their configuration over the past two millennia will have been altered substantially by human activity,
such as embankment, dredging, harbour works and lock-building, and by the influence of natural processes,
such as erosion and silt deposition. This is perfectly illustrated by the fate of the Wantsum Channel, which no
longer exists as a result of these processes.
It is possible to assess tidal behaviour in some inshore waterways, such as the Wantsum, by calculating from
modern data the hydrostatic differential between the two ends of the waterway (Grainge, 2006), although
clearly this technique cannot be applied in riverine contexts. The application to a riverine context such as the
port of London, of modern tidal data, such as the range of the tide and mean tide level, is problematic. It is
not easy to envisage the technological developments required to model the changes that would have
influenced the tidal regime in each major river in the Roman period.
Publication and dissemination
As with any other period or field of archaeology an important aspect of the research agenda is the publication
and dissemination of data and research findings. Greater encouragement needs to be given for the
publication of key research already undertaken. Moreover, greater emphasis needs to be placed on better
integration of maritime and terrestrial data and research.
In addition, many assumptions by archaeologists and historians are based, at best, on wafer-thin assessments
or, at worst, on ignorance of the characteristics of the ships of the period. Ships were built as they were for
very good reasons and clearly worked, but within limitations imposed both by the inherent performance
characteristics of the particular type and by the maritime environment in which they operated. Scholars need
to be aware of those limitations as well as the limitations of available knowledge.
Resources need also to be devoted to conservation and display. Models of the Barland’s Farm boat and St
Peter Port 1 are on display respectively at Newport Museum and Gallery and the Maritime Museum, St Peter
Port, Guernsey. Conservation and display of the original finds, though expensive would be valuable for
dissemination, as well as providing a springboard for future research.
Synthesise existing relevant data from terrestrial projects
It is a discredit to British archaeology that we have underestimated and overlooked the quality of material that
fisherman casually uncover and disregard in the offshore zone. We must address the means by which effective
archaeological liaison can be constructed and maintained with modern fishing communities who are the
natural custodians of these sites. Need to now address how the application of CHIRP and swath bathymetry to
archaeological prospection can be improved and where their use should be targeted.
The incidence of ship-losses in the final approach to ports is not uncommon so wrecks from different periods
frequently occupy virtually the same spot; the Channel approach to the Thames via the Goodwins, and the
Bembridge Ledges in the Solent approach to Portchester and Clausentum are outstanding examples. The
approaches to pre-Conquest landing places also demand archaeological prospection. The riverine approach to
Fingringhoe and Camulodunum is one example; the hazardous approach to Lulworth Cove and its attendant
hillfort on Bindon Hill is another. Some minor finds of Roman coins off Lulworth cliffs is sufficient to
demonstrate potential.
4. Maritime networks
Supply & trade – how it works, what the routes were, how these changed from C1 BC to C5 AD.
Who owned the coast?
Extent of raiding – ?Ely villa near Cardiff
Naval operations
De-settlement of east coast regions (Kent & East Anglia)
The Channel – conduit or barrier?
Points of departure on N French coast
5. Maritime identities and space
 What impact did the maritime environment and associated ‘influence’ have on communities both ‘Roman’
and ‘native’?
 How was the water itself conceived and understood?
 Who is regulating landing places, moorings and trading activities and how are these power relationships
experienced, understood and perhaps even usurped by individuals and communities?
 Although the Romans brought a measure of homogeneity across parts of Britain, the influx of new goods,
people and beliefs through maritime routes made possible more diversity of expression of identity in
local situations - how does this manifest itself in the archaeological record?
 How do ‘maritime’-focused communities which developed around ports and landing places
compare/contrast to those that developed around forts or other specialist communities?
 A shift in focus away from the technologies of functionality towards the symbolic and the ‘aesthetic’ aspects
of these vessels in order to enhance our understanding of identity?
Disconnect between Roman terrestrial and maritime archs
A maritime research agenda
It is suggested here that a fully integrated research agenda for maritime archaeology should be a priority for
England: some possible pointers for such work have been outlined above and may be summarised thus:
REGIONAL THEMES



1
2
3
plotting sea-level change over time
plotting coastal change over time, based on (1) above
development of maritime settlements in relation to each other and to (1) and (2) above
SITE-SPECIFIC THEMES


4
5
development of port topography and port buildings
development of harbours and landing places in relation to (1) (2) and (3) above
MARITIME THEMES




6
maritime industries (fishing, shipbuilding, salt etc) in relation to (1) (2) (3) (4) and (5) above
7
maritime defences, in relation to (1) (2) (3) (4) and (5) above
8
maritime transport: excavation, research and publication of wrecks, hulks and vessel
fragments
9
audit of maritime-related finds (from land and underwater sites) in the region's museums
and private collections
METHODOLOGICAL THEMES


10
development of monitoring frameworks for underwater and intertidal zone sites
11
development of surveying/ prospecting methodologies for maritime terrestrial sites and for
coastal, intertidal and fully marine environments
Such a regional maritime focus should also feed directly into any environmental, defence, urban or periodbased programmes. There is a clear need to encourage the publication of more vessel sites, be they
underwater, intertidal or 'dry-land' , as well as reviews of vessel fragments, ship fastenings, fish hooks, net
weights and other maritime-related artefacts languishing in museums or private collections, unstudied and uncatalogued: these studies should preferably be undertaken on a regional-basis. There is a surprising dearth of
harbour and shipyard excavations: perhaps some preliminary researches could be conducted to identify
possible sites before deep drainage works destroys them unremarked. As for underwater wreck sites, the
often prohibitive costs and logistical problems associated with such projects argues strongly for a
comprehensive long-term monitoring programme, rather than 'total excavation': a similar approach may will
be required for sites eroding in the inter-tidal zone.
Finally, it is worth stressing the wider relevance of underwater archaeology, and the development of new and
more efficient methodologies in that unique environment. Work below the low water mark is not just about
maritime matters, for it can also provide information on submerged palaeo-landscapes and their associated
terrestrial finds, since land surfaces pre-dating recent sea-level rises can be found offshore. There is therefore
the continuing need to integrate marine, intertidal and terrestrial surveys, to extend the 'seamless' approach
(Milne 2008).
Bibliography
Ancient Sources
Ancient sources quoted, and abbreviations used
Aurelius Victor (Aur. Vic.)
Caesares
Caesar (Caes.)
Bellum Civile (BCiv.)
Caesar (Caes.)
Bellum Gallicum (BGall.)
Dio Cassius
Roman History
Eutropius (Eutr.)
Breviarium ab urbe condita (Brev.)
Eutropius (Eutr.)
Panegyrici Latini
Pliny (the Elder)
Naturalis Historia (HN)
Strabo (Strab.)
Geographica (Geog.)
Suetonius (Suet.)
Divus Claudius (Claud.)
Tacitus (Tac.)
Agricola (Ag.)
Vegetius (Veg.)
De re militari (Mil.)
Adams, J. 2001. Ships and boats as archaeological source material. World Archaeology 32.3: 292-310.
Allason-Jones, L. 1984. A Lead Shrine from Wallsend. Britannia 15, 232-2.
Allason-Jones, L. 1996. Roman Jet in the Yorkshire Museum. Yorkshire Museum, York.
Allason-Jones, L., & Jones, J. M. 2001. Identification of ‘jet’ artefacts by reflected light microscopy. European
Journal of Archaeology 4.2: 233-251.
Allen, T. 2004. Swine, Salt and Seafood: a case study in early medieval settlement in north-east Kent.
Archaeologia Cantiana CXXIV. 117-135.
2002. The Kentish Copperas Industry. Archaeologia Cantiana CXXII. 319-333.
Allen, J.R.L & Fulford, M.G. 1992. Romano-British and later geoarchaeology at Oldbury Flats: reclamation and
settlement on the changeable coast of the Severn Estuary. Archaeological Journal 149: 82-123.
Allen, J.R.L. and Fulford, M.G. 1990. Romano-British and later reclamations on the Severn salt marshes at
Elmore, Gloucestershire. Transactions of the Bristol and Gloucestershire Archaeological Society 108: 17-32.
Allen, J.R.L. & Fulford, M.G. 1996. The distribution of South-East Dorset Black Burnished Ware Category 1
pottery in south-west Britain. Britannia 27, 223-90.
Allen, J.R.L. & Fulford, M.G. 1999. Fort Building and Military Supply along Britain’s Eastern Channel and North
Sea Coasts: The Later Second and Third Centuries. Britannia 30, 163-184.
Allen, J.R.L., Fulford, M.G. & Pearson, A.F. 2001. ‘Branodunum’ on the Saxon Shore (North Norfolk): A Local
Origin for the Building Material. Britannia 32, 271-275.
ApSimon, A.M. 1965. The Roman Temple on Brean Down, Somerset. Proceedings of the University of Bristol
Spelaeological Society 10, 195-258.
Arnold, B. 1978. Gallo-Roman boat finds in Switzerland, in Taylor & Cleere (eds.) 1978: 31-35.
Askew, G. 1980. The Coinage of Roman Britain. London.
Atkinson, M., & and Preston, S.J. 1998. The Late Iron Age and Roman settlement at Elms farm, Heybridge,
Essex, excavations 1993-5: an interim report. Britannia 29: 85-110.
Bateman, N. & A. Locker. 1982. The sauce of the Thames. The London Archaeologist 4.8 : 204-7
Beattie-Edwards, M, 1999, Studies of the possible Roman riverfront facilities at St. Deny's.
http://www.geog.port.ac.uk/webmap/itchen/overview.html
Bellamy, P.S. & G. Milne. 2003. An archaeological evaluation of the medieval shipyard facilities at Small
Hythe. Archaeologia Cantiana 123: 353-82.
Bidwell, P. 2001. A Probable Roman Shipwreck on the Herd Sand at South Shields, The Arbeia Journal 6-7: 121.
Bill, J and Clausen, B 1999. Maritime Topography and the Medieval Town. National Museum of Denmark. mono
4.
Bird, J. 1992. Samian ware, in M. Oliver, Excavation of an Iron Age and Romano-British Settlement Site at
Oakridge, Basingstoke, Hampshire 1965-66. Proceedings of the Hampshire Field Club Archaeological Society 48,
55-94.
Birley, E. 1957. Review. The Classical Review, New Series, 7(2), 173-4.
Bogaers, J.E. 1967. Einige opmerkingen over het Nedelandse gedeelte van de limes van Germania Inferior.
Germania Secunda 17, 99-114.
Bogaers, J.E. 1971. Nehalennia en de epigrafishe gegevans. Deae Nehalenniae Gids bij de tentoonstelling
Nehalennia de Zeeuwse godin, Zeeland in de Romeinse tijd, Romeinse monumentum uit de Oosterschelde, 3343.
Bogaers, J.E. & Gysseling, M. 1972. Over de naam van de godin Nehalennia. Naamkunde 4 3.4, 221-30.
Boon, G.C. 1977a. The Porth Felen anchor-stock. The International Journal of Nautical Archaeology 6.3: 23942.
1977b. A Greco-Roman anchor-stock from North Wales. The Antiquaries Journal 57: 10-30.
Brading Roman Villa http://www.bradingromanvilla.org.uk/about.php
Bradley, R. 1998. The Passage of Arms. Oxford: Oxbow Books.
Bradley, R. 1992. Roman salt production in Chichester Harbour. Rescue excavations at Chidham, West Sussex.
Britannia 23: 27-44.
Brigham, T. with Woodger, A. 2001. Roman and Medieval townhouses on the London waterfront:
excavations at Governor’s house, City of London. MoLAS Monograph 9. London.
Burnham, B.C. & Wacher, J. 1990. The Small Towns of Roman Britain. Batsford. London.
Calkin, J.B. 1953. Kimmeridge coal money and the Romano-British shale armlet industry. Proceedings of the
Dorset Natural History and Archaeological Society 75: 45-71.
Carlson, D.N. 2007. Mast-step Coins amongst the Romans. International Journal of Nautical Archaeology 36.2,
317-324.
Casey, P.J. 1994. Carusius and Allectus: The British Usurpers. London: B.T. Batsford Ltd.
Champion 2007
Cleere, H. 1977. The Classis Britannica, in D E Johnston (ed) The Saxon shore. CBA Research Report 18: 16-19.
Cleere, H. 1978. Roman harbours in Britain south of Hadrian’s Wall, in Taylor & Cleere (eds.): 36-40.
Cohen, N 2008. Churches in a Maritime Landscape: an examination of ecclesiastical activity on the Romney
Marsh. Romney Marsh Irregular. 31. 5-19.
Coles, B.P.L. and Funnel, B.M. (1981),’ Holocene palaeoenvironments of Broadland, England’. Special
Publications of the International Association of Sedimentologists (1981), 123-131.
Collingwood, R.G. and Wright, R.P. 1965. The Roman Inscriptions of Britain. Oxford.
Cook, N.C. 1971. An iron anchor, probably Roman, from Blackfriars. The Antiquaries Journal 51: 318.
Corney, M. 2002. The Roman Villa at Bradford on Avon. The Investigations of 2002. Ex Libris Press.
Cotterill, J. 1993. Saxon Raiding and the Role of the Late Roman Coastal Forts of Britain. Britannia 24, 227-239.
Craddock, P.T. & D.R. Hook. 1987. Ingots from the sea: The British Museum collection of ingots, The
International Journal of Nautical Archaeology 16.3: 201-6.
Crowson, A., Lane, T. and J. Reeve (2000), Fenland Management Project Excavations 1991-1995. Lincolnshire
Archaeology and Heritage Reports Series No. 3. Heritage Lincolnshire: Heckington.
Cunliffe, B.W. 1988.
Archaeology. Oxford.
Mount Batten: a Prehistoric and Roman Port: Oxford University Committee for
Cunliffe, B. 1972. The late Iron Age metalwork from Bulbury, Dorset. AJ 52, 293-308
Cunliffe, B. 1975. Excavations at Porchester Castle. Volume 1: Roman. London: Society of Antiquaries (Reports
of the Roman Research Committee of the Society of Antiquaries of London, XXXII).
Dark, K. & Dark, P. 1997. The Landscapes of Roman Britain. Sutton: Stroud.
Darling, M.J. & Gurney, D. 1993. Caister-on-Sea. Excavation by Charles Green 1951-55. East Anglian
Archaeological Report No. 60.
Dean, M. 1984. Evidence for possible prehistoric and Roman wrecks in British waters. The International
Journal of Nautical Archaeology 13.1: 78-9.
De Brisay, K.W. & Evans, K.A. (eds) 1975. Salt: The Study of an Ancient Industry. Colchester Archaeological
Group: Colchester.
Devoy, R.J. 1982. Analysis of the geological evidence for Holocene sea-level movements in SE England,
Proceedings of the Geologists’ Association 93: 65-90.
Devoy, R J N, 1979: ‘Flandrian sea-level changes and vegetational history of the Lower Thames Estuary’, Phil.
Trans. Roy. Soc. London Occas. Pap. 1(n. s.), 134-54? Phil. Trans. Roy. Soc. London B285, 355-407?
1977. Flandrian sea level changes in the Thames Estuary and the implications for land subsidence in England and
Wales. Nature Vol. 270: 712-5.
Diderot, D. and D’Alembert, J. (eds,) 1751-1772. L’Encyclopédie ou Dictionnaire raisonné des sciences, des arts
et des métiers, (Paris 28 vols), xii.
D’Olier, B. 1972. Subsidence and sea-level rise in the Thames Estuary. Philological Transactions of the Royal
Society London A. 272: 121-30.
Eddison, J 2000. Romney Marsh: survival on a frontier. Tempus, Stroud.
Eddison, J, Gardiner, M and Long, A 1998. Romney Marsh: environmental change and human occupation in a
coastal lowland. Oxford
Ellmers, D. 1978. Shipping on the Rhine during the Roman period: the pictorial evidence, in Taylor & Cleere
(eds.) 1978: 1-14.
1969. Keltischer Schiffbau. Jahrbuch des Römisch-Germanischen Zentralmuseums, Mainz 16: 73-122.
Esmonde Cleary, S. 1989. The Ending of Roman Britain. London: Batsford.
Fawn, A. J. et al. 1990. The Red Hills of Essex : Salt-making in Antiquity. Colchester Archaeological Group:
Colchester.
Fincham, G. 2002. Landscapes of Imperialism: Roman and Native Interaction in the East Anglian Fenland.
British Archaeological Reports British Series 338. Archaeopress: Oxford.
Firth, A. 1998. Shoreline management plans strategic study. Wessex Archaeology.
Fitzpatrick, A.P. 1984. The deposition of La Tene Iron Age metalwork in watery contexts in southern England, in
B.W. Cunliffe and D. Miles (eds.) Aspects of the Iron Age in Central Southern Britain. Oxford University
Committee for Archaeology Monograph 2, 178-90.
Fryer, J. 1973. The harbour installations of Roman Britain, in Blackman, D. (ed.) Marine Archaeology. London:
Butterworths. 261-76.
Fulford, M. 2001. Links with the Past: Pervasive ‘Ritual’ Behaviour in Roman Britain. Britannia 32, 199-218.
Fulford, M. 1989. A Roman Shipwreck off Nornour, Isles of Scilly? Britannia 20: 245-9.
1984. Demonstrating Britannia’s Economic Dependence in the First and Second Centuries, in Blagg, T.
& King, A. (eds.) Military and Civilian in Roman Britain: Cultural Relationships in a Frontier Province.
BAR British Series 136: 129-41.
1978. The interpretation of Britain’s late Roman trade: the scope of medieval historical and
archaeological analogy, in Taylor & Cleere (eds.) 1978: 59-69.
Fulford, M.G. & Allen, J.R.L. 1992. Iron-making at the Chesters Roman Villa, Woolaston. Britannia 23: 159–
215.
Fulford, M.G., Allen, J.R.L. and Rippon, S. (1994), The settlement and drainage of the Wentlooge Level, Gwent:
Survey and excavation at Rumney Great Wharf, 1992. Britannia XXV: 175-211.
Fulford, M., Champion, T. & Long, A. 1997. England’s Coastal Heritage: A survey for English Heritage and
RCHME. London: English Heritage and RCHME.
Fulford, M. & Tyers, I. 1995. The date of Pevensey and the defence of an ‘Imperium Britanniarum.’ Antiquity
69, 1009-14.
Galliou, P. 1982. Les amphores tardo-républicaines découvertes dans l’ouest de la France et les importations
de vin italien à la fin de l’Age de Fer. Fasc. I du Corpus des amphores découvertes dans l’ouest de la France, dir.
Sanquer, R. Archaeologie en Bretagne, Suppl. 4, Brest.
Gardiner, A. 2007. An Archaeology of Identity: Soldiers and Society in Late Roman Britain. Walnut Creek, CA:
Left Coast Press.
Grainge, G. 2006. Double tides in the Wantsum: fact or fiction? Archaeologia Cantiana. CXXVI. 381-91.
Grainge, G. 2002. The Roman Channel Crossing of A.D. 43: The constraints on Claudius’s naval strategy.
Oxford: BAR British Series 332
Grant, A. 1975. The Animal Bone, in B.W. Cunliffe (ed.) Excavations at Porchester Castle. Volume 1: Roman.
London: Society of Antiquaries (Reports on the Research Committee of the Society of Antiquaries of London,
XXXII), 378-408.
Good, C. & Plouviez, J. 2007. The Archaeology of the Suffolk Coast. Suffolk County Council Archaeological
Services Report.
Gurney, D. 1986. Settlement, religion and industry on the Fen-edge: three Romano-British sites in Norfolk.
East Anglian Archaeology Report 31. Dereham: Norfolk Archaeology Unit.
Hall, D. and Coles, J.M. 1994. Fenland Survey. An Essay in Landscape and Persistence. London: English
Heritage.
Hardman, F. W. & Stebbings, W. P. D. 1940-1942. Stonar and the Wantsum Channel, (parts 1-3). Archaeologia
Cantiana 53-55, 62-80; 41-55 & 37-49.
Harmand, L. 1966. A propos d’un col d’amphore trouve dans la Manche, in Heurgon, J. et al. (eds.) Melanges
d’Archaeologie, d’Epigraphie et d’Histoire Offerts a Jerome Carcopino. pp. 477-89. Paris.
Hartley, K. 1977. Two major industries producing mortaria in the first century AD, in J Dore and K Greene
(eds.) Roman pottery studies in Britain and beyond. BAR 30:5-17.
Hassall, M. 1978. Britain and the Rhine provinces: epigraphic evidence for Roman trade, in J. du Plat Taylor
and H. Cleere (eds.) Roman Shipping and Trade: Britain and the Rhine Provinces. London: CBA Research Report
24: 41-8.
Henig, M. and Ross, A. 1998. A Roman Intaglio Depicting a Warship from the Foreshore at King’s Reach,
Winchester Wharf, Southwark. Britannia, 29: 325-327.
Hinchcliffe, J. & Sparey Green, C. (eds.) 1985. Excavations at Brancaster 1974 and 1977. East Anglian
Archaeology 23.
Hingley, R. 2000. Roman Officers and English Gentlemen: The Imperial Origins of Roman Archaeology. London:
Routledge.
Hingley, R. 2006. The Deposition of Iron Objects in Britain during the Later Prehistoric and Roman Periods:
Contextual Analysis and the Significance of Iron. Britannia 37, 213-267.
Hoare, P et al 2002. Reused Bedrock ballast in King's Lynn Town Walls. Medieval Archaeology 46. 91-105.
Höckmann, O. 1986. Römische Schiffsverbände auf dem Ober- und Mittelrhein und die Verteidigung der
Rheingrenze in der Spätantike, Jahrbuch des Romanisch-Germanischen Zentral Museums Mainz 33: 369-415.
Hondius-Crone, A. 1955. The Temple of Nehalennia at Domburg. Amsterdam: J.M. Meulenhoff.
Hurst, H.R. 1999. Civic space at Glevum, in Hurst, H.R. (ed.) The coloniae of Roman Britain: new studies and a
review. Journal of Roman Archaeology. Portsmouth pp. 113–25.
Ingold, T. 1993. The Temporality of the landscape. World Archaeology 25 (2) 24-174.
Johnson, M. 1999. Archaeological Theory: An Introduction. Oxford: Blackwells.
Johnson, S. 1976. The Forts of the Saxon Shore. London: Book Club Associates.
Johnston, D.E. 1977 (ed.) The Saxon Shore. Council for British Archaeology Research Report 18.
Jones, J. Ellis 2009. The maritime and Riverine Landscape of the West of Roman Britain: Water Transport on
the Atlantic Coasts and Rivers of Britannia. British Archaeological Report 493. Archaeopress. Oxford.
Jones, G.D.B. & Mattingly, D. 1990. An Atlas of Roman Britain. Blackwell. Oxford.
Jones, M.J. 2002. Roman Lincoln: Conquest, Colony and Capital. Tempus. Stroud.
Kiernan, P. 2009. Miniature Votive Offerings in the North-West Provinces of the Roman Empire. Wiesbaden:
Verlag.
King, A. & Soffe, G. 1994, The Iron Age and Roman temple on Hayling Island. In A P Fitzpatrick & E L Morris
(eds) The Iron Age in Wessex: Recent Work. Salisbury: Wessex Archaeology, 114 -16.
L’Hour, M. 1987. Un site sous-marie sur la cote de L’Armorique l’epave antique de Ploumanac’h, in Revue
Archaeologique de L’Ouest: 113-131.
Locker, A. 2007. In piscibus diversis: the bone evidence for fish consumption in Roman Britain. Britannia 38:
141–80.
Long, A, Hipkin, S & Clark, H (eds.) 2002. Romney Marsh: coastal and landscape change through the ages.
OUCA monograph 56.
Long, A.J. and Roberts, D.H. (1997), ‘Sea level change’, in Fulford, M., Champion, T. and Long, A. England’s
Coastal Heritage. A survey for English Heritage and the RCHME. English Heritage Archaeological Report 16,
25-49. London: English Heritage.
Lydden Valley Research Group, (2006), The Geology, Archaeology and History of Lydden Valley and Sandwich
Bay. Dover.
Lyne, M. 1999. Roman ships’ fittings from Richborough. Journal of Roman Military Equipment Studies 7: 1479.
Markey , M. 1997. An inscribed stone anchor from Dorset, The International Journal of Nautical Archaeology
26.2: 127-32.
1991. Two stone anchors from Dorset, The International Journal of Nautical Archaeology 20.1: 47-51.
Marsden, P. 1994. Ships of the Port of London: First to eleventh centuries AD. English Heritage:
Archaeological Report 3, London.
1990. A re-assessment of Blackfriars Ship 1, in McGrail (ed) Maritime Celts, Frisians and Saxons. CBA
Research Report 71: 66-74.
1977. Celtic Ships of Europe, in S. McGrail (ed.) Sources and Techniques in Boat Archaeology. BAR 29:
281-8.
1967. A Boat of the Roman Period Discovered on the Site of New Guy's House, Bermondsey, 1958,
Transactions of the London and Middlesex Archaeological Society 21: 118-31.
Martin, C. 1992. Water Transport and the Roman Occupations of North Britain, in Smout, T.C. (ed.) Scotland
and the Sea. Edinburgh: John Donald Publishers pp. 1–34.
Mason, D.J.P. 2003. Roman Britain and the Roman Navy. Tempus.
https://shareweb.kent.gov.uk/Documents/Leisure-and-culture/heritage/serf-seminar-notes/maritime.pdf
McDonald, K. 1977. The Treasure Divers.
McGrail, S. 1995. Romano-Celtic Boats and Ships: Characteristic Features. The International Journal of Nautical
Archaeology 24.?: 139-45.
McGrail, S. 1987. Ancient Boats in N W Europe: The Archaeology of Water Transport to AD 1500. London &
New York: Longman.
McGrail, S. 1983. Cross-Channel Seamanship and Navigation in the Late First Millennium BC. Oxford Journal of
Archaeology 2.3: 299-337.
McGrail, S., & Roberts, O. 1999. A Romano-British Boat from the Shores of the Severn Estuary, Mariner's
Mirror 85: 133-46.
McInnes, R. (2009), Coastal Risk Management: a Non-Technical Guide. Ventnor: Standing Conference on
Problems Associated with the Coast (SCOPAC).
Merrifield, R. 1962. Coins from the bed of the Walbrook and their significance. Antiquaries Journal 42, 38-52.
Merrifield, R. 1987. The Archaeology of Ritual and Magic. London: Guild Publishing.
Middleton, P. 1983. The Roman Army and Long Distance Trade, in Garnsey & Whittaker (eds.) 1983: 75-83.
Millett, M. 1990. The Romanization of Britain: An Essay in Archaeological Interpretation. (Reprinted 1994)
Cambridge: Cambridge University Press.
Millet, M. 2007. Roman Kent, in Williams, J.H. [ed] The Archaeology of Kent to AD800. Boydell.
Milne, G. 2008. Maritime Archaeology. South-East Research Framework
2007. Notes on the South East Research Framework Maritime Themes,
2003. The Port of medieval London. Tempus.
2001. Joining the Medieval Fleet. British Archaeology 61. CBA.
2000. A Roman Provincial Fleet: the Classis Britannia reconsidered in Oliver, G., R Brock, T Cornell & S
Hodkinson (eds) The Sea in Antiquity, 127-31.
1995. Roman London. English Heritage. London.
1985. The Port of Roman London. London: Batsford.
MoLAS, 2007, Uncovering the Past: Archaeological discoveries in Chichester Harbour AONB 2004 – 2007.
Chichester Harbour Conservancy.
Monaghan, J. 1991. Pottery from marine sites around Guernsey. J Roman Pottery Studies, 3, 63-9.
1991? Decouvertes maritimes provenant du baillage du Guernsey, SFECAG Actes du Congres du Caen
39-44
1989 ‘The Guernsey Maritime Trust Gazetteer 1985-88’, TSG XXII, III, 453-65
Moore, J. & Heathcote, J. (Spring) 2004. Geochemistry: Richborough. Conservation Bulletin 45. English
Heritage, 12-13.
Morris, J. 1982. Londinium: London in the Roman Empire. London: Weidenfeld & Nicolson.
Morrison, J.S., & Coates, J.F. 1996. Greek and Roman Oared Warship. Oxford: Oxbow Books.
Morrison, J.S., J.F. Coates, & N.B. Rankov. 2000. The Athenian Trireme: the History and Reconstruction of an
Ancient Greek Warship (2nd ed). Cambridge, New York and Oakleigh: Cambridge University Press.
Muckelroy, K. 1978. Maritime Archaeology. Cambridge: Cambridge University Press.
Muckelroy, K., C. Haselgrove & D. Nash. 1978. A pre-Roman coin from Canterbury and the ship represented on
it. Proceedings of the Prehistoric Society 44: 439-44.
Murphy, P. (2005), Coastal change and human response, in Ashwin, T. and Davison, A.(eds.) An Historical
Atlas of Norfolk (3rd edition),.pp. 6-7. Phillimore: Chichester.
Naish, J 1985. Seamarks: their history and development. London, Stanford Maritime.
NAVIS I & NAVIS II database http://www2.rgzm.de/navis/home/frames.htm
Nayling, N., & S. McGrail. 2004. The Barland’s Farm Romano-Celtic boat. York
Nayling, N., D. Maynard, & S. McGrail. 1994. Barland’s Farm, Magor, Gwent: A Romano-Celtic Boat. Antiquity
68 : 596-603.
Orna-Ornstein, J., 1995 Money for Rome's naval secrets. British Archaeology, no 8, October 1995
Osborne, R. 2004. Hoards, Votives, Offerings: The Archaeology of the Dedicated Object. World Archaeology
26(1), The Object of Dedication, 1-10.
Ottaway, P. 1993. Roman York. Batsford/English Heritage. London.
Parker Pearson, M. 2003. The British and European Context of Fiskerton, in N. Field and M. Parker Pearson,
Fiskerton: An Iron Age Timber Causeway with Iron Age and Roman Offerings. Oxford: Oxbow Books, 179-89.
Pearson, A. 2002a. The Roman Shore Forts: Coastal Defences of Southern Britain. Stroud: Tempus.
Pearson, A. 2002b. Construction of the Saxon Shore Forts. BAR British Series ???.
Pearson, A. 2005. Barbarian Piracy and the Saxon Shore: A Reappraisal. Oxford Journal of Archaeology 24(1),
73-88.
Pearson, A. 2006. Piracy in Late Roman Britain: A Perspective from the Viking Age. Britannia 36, 337-353.
Perkins, D 2006. Prehistoric Maritime Traffic in the Dover Straits and Wantsum. Archaeologia Cantiana. CXXVI.
279-93.
Petts, D. 1998 ‘Landscape and Cultural Identity in Roman Britain’ In XXXXX Cultural identity in the Roman
Empire, Routledge.
Petts, D & Gerrard, C. 2006. Shared Visions: The North-East Regional Research Framework for the Historic
Environment. Durham County Council.
Phillips, C. W. (ed.) 1970. The Fenland in Roman Times: Studies of a Major Area of Peasant Colonization with a
Gazetteer Covering Known Sites and Finds. Royal Geographical Society Research Series 5, London.
Philp, B. 1996. The Roman Fort at Reculver. Kent Archaeological Rescue Unit (9th Edition).
Philp, B. 1981. The excavation of the Roman forts of the Classis Britannica at Dover, 1970-1977. KARU
Monograph 9.
Philp, B.J. 1981. The excavation of the Classis Britannica fort at Dover 1970-1977. Kent Monograph Series 3.
Philp, B. 2005. The Excavation of the Roman Fort at Reculver, Kent. Dover: Kent Archaeological Rescue Unit.
Pitassi, M. 2009. The Navies of Rome. Boydell.
Potter, T. W. 1981. The Roman Occupation of the Central Fenland. Britannia 12: 79-134.
Pownall, T. 1778. Memoire on the Roman Earthenware fished up within the Mouth of the River Thames,
Archaeologia 5: 282-90.
Reilly, J. C. 1975 Ships of the Unitied States Navy: Chistening, Launching and Commissioning. Naval History
Division: Washington.
Revell, L. 2009 Roman Imperialism and Local Identities. Cambridge University Press.
Riddler, I. 1998. Worked Whale Vertebrae. Archaeologia Cantiana CXVIII. 205-215.
Ridgeway, V 2000. A Medieval Saltern Mound at Millfield, Bramber, West Sussex. Sussex Archaeological
Collections. 138. 135-52.
Riley, W, and Gomme, L. 1912. Ship of the Roman Period discovered on the site of the new County Hall.
London.
Rippon, S. 2008 Coastal trade in Roman Britain: the investigation of Crandon bridge, Somerset, a RomanoBritish trans-shipment port beside the Severn estuary. Britannia 39: 89-144.
Rippon, S. 2000. The Transformation of Coastal Wetlands. Exploitation and management of Marshland
Landscapes in North-West Europe during the Roman and Medieval Periods. Oxford: Oxford University Press.
Rule, M. & Monaghan, J. 1993. A Gallo-Roman Trading Vessel from Guernsey: the excavation and recovery of
a third century shipwreck. Guernsey Museum Monologue No. 5.
Salway, P. 1981. Roman Britain. Oxford: Clarendon Press.
Sealey, P.R. & P.A. Tyers. 1989. Olives from Roman Spain: A Unique Amphora Find in British Waters, The
Antiquaries Journal 69.1: 53-72.
Severin, T. 1978. The Brendan Voyage. London: Hutchinson (republished by Arrow Books 1988)
Shrubsole, G.W. 1887. The Traffic between Deva and the Coast of North Wales in Roman Times.
Chester Archaeological and Historical Society Journal, N.S. 1: 71–113.
Sidell, E.J., Wilkinson, K.N., Scaife, R.G. and N. Cameron, (2000), The Holocene Evolution of the River Thames.
Archaeological Investigations (1991-1998) in Advance of the London Underground Limited Jubilee Line
Extension. Museum of London Archaeology Service Monograph 5. London: MOLAS.
Simmons, B.B. (1980), ‘Iron Age and Roman coasts around the Wash’, in Thompson, F.H. (ed.), Archaeology
and Coastal Change, London: Society of Antiquaries, pp. 56-73.
Sleightholme, J.D. 1970. Cruising: a Manual for the Small Sailing Boat Owner (2nd ed). London: Adlard Coles
Spaul, J. 2002. Classes Imperii Romani. Nectoreca Press.
Spurrell, F.C.J. 1885. Early Sites and Embankments on the Margins of the Thames Estuary, The Archaeological
Journal 42: 269-302.
Starr, C. 1993. The Roman Imperial Navy 31 BC – AD 324 (3rd ed). Ares Publishers.
Steers, J. A. 1964. The coastline of England and Wales. Cambridge.
Steane, J and Foreman, M 1991. The archaeology of medieval fishing tackle. In G Good et al (eds.) Waterfront
Archaeology. CBA Res Rep 74. 88-101.
Stuart, P. & Bogaers, J.E. 1971. Catalogues van de monumentum. Deae Nehalenniae Gids bij de tentoonstelling
Nehalennia de Zeeuwse godin, Zeeland in de Romeinse tijd, Romeinse monumentum uit de Oosterschelde, 3343.
Swain, H. 1986. A note on the Romano-British and other finds from Carr House Sands, Seaton Carew,
Cleveland. Northern Archaeology 7.2: 31-34.
Taylor, J.d.P. & H. Cleere. (eds.) 1978. Roman shipping and trade: Britain and the Rhine provinces. CBA
Research Report 24.
Todd, M. 1992. Jet in Northern Gaul. Britannia 23: 246-248.
Toft, L.A. (1992), ‘Roman quays and tide levels’. Britannia 23, 249-254.
Tomalin, D.J. 2006. Coastal villas, maritime villas; a perspective from southern England. Journal of Maritime
Archaeology 1.1: 29-84
Tomalin, D.J. 1997. Bargaining with nature, considering the sustainability of archaeological sites in the dynamic
environment of the inter-tidal zone. Proceedings of the Conference of 1st-3rd April 1996 at the museum of London;
preservation of archaeological remains in situ. Museum of London/University of Bradford. 144-158.
Tomlin, 1997. Inscriptions. Britannia 27: 455 - 458
Topping, P and Swan, V 1995. Early Salt-working sites in the Thames Estuary. In RCHME 1995. 28-40.
Waller, M. (1994), The Fenland Project No. 9: Flandrian Change in Fenland. East Anglian Archaeology 70.
Cambridge: Fenland Project Committee/Cambridgeshire County Council.
Walsh, M. 2006. Pudding Pan: A Roman shipwreck and its cargo in context. (Unpublished) Thesis submitted
for the degree of Doctor of Philosophy at the University of Southampton in 2006. University of
Southampton.
2002. Pudding Pan: A case study on the enhancement of the Romano-British maritime record, in Muskett,
G. & M. Georgiadis (eds.) The Seas in Antiquity: 76-87. Liverpool: University of Liverpool.
1998. The Riddle of the Sands: An Assessment of the Potential for Identifying Roman Shipwrecks in British
Waters. (Unpublished) Dissertation submitted in partial fulfilment for the degree of MA (Maritime
Archaeology) of the University of Southampton in 1998. University of Southampton.
Webster, J. 1997. ‘Text expectations: the archaeology of ‘Celtic’ ritual wells and shafts, in A. Gwilt & C.C.
Haselgrove (eds.). Reconstructing Iron Age Societies. Oxbow Monograph 71, 216-27.
Weerd, M. D. de, 1988. A Landlubber’s View of Shipbuilding Procedure in the Celtic Barges of Zwammerdam,
The Netherlands, in Lixa Filgueiras, O. (ed.) 1988. Local Boats: Fourth International Symposium on
Boat and Ship Archaeology, Porto 1985. British Archaeological Series International Series 438 (i): 3552.
1978. Ships of the Roman period at Zwammerdam/Nigrum Pullum, Germania Inferior, in Taylor &
Cleere (eds.) 1978: 15-21.
Willis, S. 1998. Samian Pottery in Britain: Exploring its distribution and archaeological potential. The
Archaeological Journal 155, 82-133.
Willis, S. 2005. Samian Pottery, a Resource for the Study of Roman Britain and Beyond: the results of the
English Heritage funded Samian Project. Internet Archaeology 17.
Wilmott, T. & Tibber, J. 2009. Richborough, a Roman and Medieval Port. Research News 12. English Heritage,
20-22.
Wilson, P.R. (1989), ‘Aspects of the Yorkshire Signal Stations’, in Maxfield, V.A. and Dobson, M.J. (eds), Roman
Frontier Studies 1989. Proceedings of the XVth International Congress of Roman Frontier Studies. Exeter:
University of Exeter Press, pp. 142-7.