[go: up one dir, main page]

Academia.eduAcademia.edu
PAPUA NEW GUINEA LIQUEFIED NATURAL GAS PROJECT SOCIAL IMPACT ASSESSMENT 2008 Assoc Professor Laurence Goldman January 2009 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5 Portion 152 4.5.1 Introduction The Papua New Guinea (PNG) Liquefied Natural Gas (LNG) Project (‘the project’ or ‘PNG LNG Project’) aims to commercialise the gas resources in the Southern Highlands and Western provinces of PNG. The project involves the production of gas and its transportation to an LNG plant at Portion 152 on the coast of the Gulf of Papua near Port Moresby. The gas is to be liquefied at the LNG plant and the LNG product shipped to international markets. The PNG LNG Project will be operated by an ExxonMobil Corporation (ExxonMobil) affiliate, Esso Highlands Limited (Esso). This study has been commissioned by Coffey Natural Systems (CNS), the Environmental Impact Study (EIS) lead consultants for the project. This report documents cultural heritage sites found within and near 1) the proposed Portion 152 LNG Facilities site (referred to throughout this report as the ‘Onshore LNG Facilities component’ of our surveys); and 2) the proposed associated shallow marine area of Caution Bay, including the inter-tidal zone (referred to as the ‘Offshore LNG Facilities component’) (Figure 4.5.1). Combined, these two components are referred to as ‘the Study Area’ (see section 4.5.3.1 for exact location of the Study Area). This report begins with an outline of the proposed LNG Facilities developments in the Study Area in section 4.5.2 (as these have been communicated to us by Coffey Natural Systems), and the objectives of the cultural heritage sites surveys in the ensuing section 4.5.3. Legislative, ethical and best practice issues that guide the cultural heritage surveys presented in this report are outlined in section 4.5.4. This is followed by the methodology employed to document the Study Area’s cultural heritage sites (section 4.5.5). In section 4.5.6 background environmental details are presented, followed in section 4.5.7, section 4.5.8 and section 4.5.9 by cultural, archaeological and historical details relevant to understanding cultural heritage site locations and locational patterns within the Portion 152 area. This is followed by section 4.5.10, where the results of the field surveys are presented. The significance of the Study Area’s cultural heritage sites are then assessed (section 4.5.12), so as to make informed recommendations (section 4.5.14) for their management in light of potential impacts from the proposed developments (section 4.5.13). For this report, cultural heritage sites are defined as places that relate to the cultural history of the study region, as they concern the peoples engaged with them in the past and/or in the present. While this report focuses on specific places, it recognises also that the meaning of what constitutes a cultural heritage site varies widely and can include places of archaeological (scientific), historical (social) and/or traditional (cultural) values. The cultural heritage sites in this report are therefore considered in the wider context of the social landscape setting in which they occur. B. David, B. Duncan, M. Leavesley 4-451 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.2 Project Description The following descriptions of the planned LNG Facilities developments are taken directly from Attachment A of the cultural heritage consultancy agreement of 18 December 2007 between Coffey Natural Systems and Monash University, as supplied by Coffey Natural Systems. The proposed development aims to move compressed hydrocarbon gas from existing gas and oil fields through a submarine pipeline to processing and storage facilities at Portion 152. At Portion 152, the plant will receive the pipeline at inlet facilities. Processing will here involve acid-gas removal and disposal, gas dehydration involving water and mercury removal, gas liquefaction, LNG storage and loading in storage tanks, condensate storage and loading in tanks, and possible future Liquified Petroleum Gas [LPG] storage. To perform these tasks, the Portion 152 processing plant will include a range of utilities including power generation facilities, fuel systems, heating and cooling systems, nitrogen production systems, fresh water and potable water systems, and instrument air systems. The LNG Facilities will require refrigerant storage and make-up system, wet and process flares and liquid blowdown, LNG storage flare, firewater supply wells, reserve storage, distribution systems, and diesel fuel systems. Infrastructure at the LNG Facilities site will include a temporary camp for construction and contract personnel of approximately 7500 people; an operations camp for operating and maintenance personnel (approximately 500 people); waste and effluent collection/treatment systems; administration and maintenance buildings associated with the LNG Facilities; a helipad; a possible project air strip; a loading terminal; and upgrade of the existing road between the Facility and Port Moresby. A conventional piled access trestle offshore and earthen causeway across the salt pan, with a total length of 2 km, is proposed to connect the incoming marine pipeline with the onshore LNG processing and storage Facility. The trestle and causeway will support a range of facilities including the product loading platforms at the end of the trestle in 14-15 m of water depth at lowest astronomical tide (LAT). The loading platform will consist of LNG loading and return vapour arms and LNG berthing facilities. Loading platforms will be sited adjacent to the trestle, and in future possibly adjacent to the LNG or condensate loading platform. A separate materials offloading facility (MOF) is proposed to facilitate the construction of the LNG Facilities. The offloading dock is an approximately 500 m long earthen causeway with a sheet-piled dock in 7 m of water depth at LAT, capable of receiving 5000 tonne barges. LNG tankers are 3 anticipated to have a capacity of 210,000 m and require approximately 15 m water depth at LAT. LPG 3 tankers are anticipated to have a capacity up to 85,000m and require approximately 15 m water depth at LAT. Condensate tankers are anticipated to require approximately 8 m water depth at LAT. The approach and departure manoeuvres of all the carriers are expected to be assisted by 80 tonne tugs. The LNG Facilities site will be fenced with regular security inspections. B. David, B. Duncan, M. Leavesley 4-453 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.2 Location of Survey Focus Area (purple), original kidney-shaped area (blue), site security fence area (black), within and near Portion 152 (red) (as supplied by Coffey Natural Systems). The proposed LNG Facilities developments will undoubtedly require the construction of numerous access roads and tracks, laydown areas, cleared buffer zones and other infrastructure in addition to the above described processing facilities. Individually and together, these developments will certainly create considerable disturbance (e.g. through crushing or direct removal) to surface and sub-surface cultural heritage sites, both through direct disturbance (e.g. digging activity) and indirect disturbance (e.g. the passing of heavy machinery over archaeological deposits; foot traffic; access to nearby archaeological sites and the picking up of archaeological objects by personnel; increased runoff which may increase erosion along Ruisasi Creek – as the North Vaihua River is known locally based on preliminary information obtained during fieldwork for this report – and Vaihua River). Potential spillages and longer-term extensions of facilities will also almost certainly cause incremental impacts on cultural sites both within and surrounding the Study Area. B. David, B. Duncan, M. Leavesley 4-454 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.3 Proposed offshore LNG Facilities option locations in Caution Bay, showing areas where maritime cultural heritage surveys have been undertaken (within the blue 400 m and green 700 m buffer zones) (as supplied by Coffey Natural Systems). Cultural heritage sites will certainly be impacted by these planned construction, infrastructural and operational developments. In order to document the impacts of the planned developments on cultural heritage sites, a sample of the Study Area was systematically surveyed to allow probabilistic extrapolations to be made for the entire area for purposes of the Environmental Impact Study (EIS). While this report thus documents in detail the cultural heritage sites of the Survey Focus Area (a delimited, gridded area within the proposed LNG Facilities’s site security fence area; see Figure 4.5.2 and section 4.5.3.1 for definition of ‘Survey Focus Area’) and nearby transect areas and associated offshore facilities (Figure 4.5.3), at the time of this study field access to other parts of the proposed LNG Facilities area was restricted due to the unexpected discovery of unexploded ordnances (UXO). Therefore the location of cultural heritage sites outside the field-surveyed areas cannot be directly or systematically addressed in this report, although the results of the surveys can be used to undertake general predictive modeling for the LNG Facilities as a whole. This means that the specific locations of any potential infrastructure locations associated with the proposed LNG Facilities developments, including access roads, laydown areas and the like, are not addressed in this report. Such areas will require cultural heritage investigation and management planning prior to development (see section 4.5.14 for recommendations). 4.5.3 Objectives The aims of this report are to 1) document the location and nature of cultural heritage sites within the Survey Focus Area of the proposed LNG Facilities site (Figure 4.5.2) and the adjacent near-shore area (Figure 4.5.3); 2) identify potential impacts of proposed developments on these cultural heritage sites; 3) identify the significance of these cultural heritage sites; and 4) make recommendations on how to manage the area’s cultural heritage sites based on anticipated impacts and significance assessments. B. David, B. Duncan, M. Leavesley 4-455 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project The brief for the cultural heritage site surveys expects the consultants ‘to liaise with PNG national agencies in a manner that assists capacity building and national database records’, ensuring that the work meets ‘the requirements of PNG legislation and the Equator Principles. In particular the assessment should conform to IFC Performance Standard 8, Cultural Heritage’. It requires performance of the following tasks • Attend study briefing meeting with Coffey Natural Systems project staff, either in person or via teleconference: – To present an overview of the project scope and components. – To discuss and agree the study scope of work. – To discuss and agree study information requirements. – To discuss study in field travel and survey OHS, logistics and resourcing requirements. – To agree format and timing of study deliverables. • Literature review of archaeological work in the relevant project environs to date. • Inspect PNG National Museum and Art Gallery register of archaeological sites in the project area. • Liaison with Project Social Mapping and Land Identification researchers as required on information to support this study. • Provide input to planning of study field logistics and engagement of research assistants as required. • Undertake fieldwork to identify and assess the significance of archaeological and cultural heritage sites in the project area, to include: – Foot-surveys of the land component to identify archaeological and cultural heritage sites in the project area. – Informant interviews concerning significant sites (including information on landowner attitudes and expectations regarding the management of the sites). – Marine archaeological survey in near-shore area including a helicopter fly-over (if required), remote sensing survey and underwater archaeological inspection. – Photos and accurate GIS data of identified sites (including marine sites). • Liaise with the Contracting Party, project design engineers and the Social Impact Assessment (SIA) consultant Laurence Goldman on the findings of the survey to agree appropriate mitigation measures to manage potential impacts to sites or constraints to route planning. Any such measures recommended by Monash University may be reported or disclosed to third parties, including PNG Government agencies, provided that the Contracting Party (Coffey Natural Systems) provides its consent prior to the disclosure or report (such consent not to be withheld unreasonably). • Consultation with and presentation of the findings of the surveys to the PNG National Museum and Art Gallery and the Department of Environment and Conservation (DEC), as appropriate. • Formal presentation to DEC of the findings of the study as part of the EIS process. • Produce a report addressing the objectives listed above. The present report fulfils this task. B. David, B. Duncan, M. Leavesley 4-456 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project We describe below under two separate sections – ‘LNG Land’ and ‘LNG Shallow Marine’, representing the two separate geographical and methodological components of this study – specific locational details of the Study Area. 4.5.3.1 Onshore LNG Facilities Component The Survey Focus Area which forms the focus of this study is located between Papa and Boera villages (see Figure 4.5.2). The location and size of the land area to be covered by the present study changed during the course of the cultural heritage surveys. Initially the Monash University cultural 2 heritage site surveying team was asked to cover an area of land 8.18 km in size. However, the discovery of an UXO early during the cultural heritage surveys (see section 4.5.5.1) resulted, at the request of Coffey Natural Systems, in a cessation of the original surveys and changes to the brief to allow UXO clearance by a technical team of experts prior to continuation of the cultural heritage field surveys. This led Coffey Natural Systems to refine the survey area to what is referred to in this report as the Survey Focus Area throughout, consisting of an area 2.2 km (east-west) x 1.2 km (north-south) in size as shown on Figure 4.5.2, and representing part of the focal area of construction activity for the proposed LNG Plant. No planned access roads, laydown areas or other infrastructure places outside the Survey Focus Area were investigated during our surveys, although some cultural heritage sites outside the Survey Focus Area were opportunistically recorded and are documented in this report. Figure 4.5.4 The kidney-shaped area (blue line) within and near Portion 152 (red line) for which cultural heritage sites surveys (land component) were originally requested (as supplied by Coffey Natural Systems). B. David, B. Duncan, M. Leavesley 4-457 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project The cultural heritage sites documented for this report are located within and near the PNG Government-owned Portion 152, in Central Province. Portion 152 extends from latitude 09° 19' 26'' in the north (~2 km south of Papa on the 1:50,000 Port Moresby Sheet 8379-III [Edition 1] Series T702 topographic map) to 09° 23' 53'' in the south (~1 km north of Boera). Its western boundary is variously between 400 m and 800 m inland from the high water line as per government regulations. The eastern boundary runs from Kokoro Hill (~8 km east and inland from the coast) south along the foothills of the range to a point ~500 m north of the Boera turn-off, from where it runs basically on an east-west axis between the foothills and the coast. The kidney-shaped area originally marked for archaeological investigation (Figure 4.5.4), and the subsequent Survey Focus Area that superceded it and that largely lies within it (Figure 4.5.2), are located in the northwestern corner of Portion 152. Because of the above fieldwork history, the cultural heritage field surveys were undertaken over two field trips (section 4.5.3.1 for details). The result is 100% surface surveying of the northwestern and western parts of the Survey Focus Area, and transect (probabilistic) sampling for the rest of the Survey Focus Area. An area outside and to the northwest of the Survey Focus Area was also 100% surveyed for cultural heritage sites during the first field trip prior to the discovery of the UXO, when the original Study Area was larger than the Survey Focus Area itself and extended into this locality. The exact locations surveyed, methods employed, and results of the surveys are fully described in section 4.5.3.1, section 4.5.1 and section 4.5.10 respectively. 4.5.3.2 Offshore LNG Facilities Component Our brief for the Offshore LNG Facilities component of the cultural heritage site surveys was to 2 undertake a maritime archaeological survey of a ~4km area of Option One (with a 400m buffer zone around proposed locations for an LNG Jetty) as a priority, with further investigations of a 700m buffer zone (Option 2) around these same proposed locations if any time remained available during our period of contract. The shallow marine area covered by this report is shown in Figure 4.5.3. Here the area surveyed for cultural heritage sites includes that part of the route which would be directly impacted by the construction of the proposed offshore LNG Jetty near Konekaru beach and the proposed Marine Offshore Facility near Boera Head (as determined by the map supplied to the cultural heritage site surveying team by Coffey Natural Systems immediately prior to fieldwork – see Figure 4.5.3). Documentation of cultural heritage sites in other nearby underwater areas did not form part of the brief, budget, schedule or logistical planning; hence any inspection of such additional areas was opportunistic and restricted. It should be noted that shallow marine areas outside the Study Area addressed in this report will require cultural heritage investigation and management planning prior to development if such areas are to be impacted by the proposed developments. 4.5.4 Law, Ethics and Best Practice Current international best practice standards for extractive industries are laid out in the World Bank/IFC-sponsored Equator Principles (Equator Principles, 2005), and the legislative requirements have been detailed in Section 4.1. However, we reiterate here those sections that are relevant to Portion 152. Current international best practice standards for extractive industries are laid out in the World Bank/International Finance Corporation (IFC)-sponsored Equator Principles (Equator Principles, 2006). The IFC standards required under the Equator Principles incorporate IFC Performance Standard 8 (Cultural Heritage), which is of relevance to development responsibilities towards a variety of archaeological and non-archaeological cultural heritage places documented in this report. IFC Performance Standard 8 adopts the United Nations Educational, Scientific and Cultural Organization’s (UNESCO) definition of ‘cultural property’ which includes: sites having archaeological (prehistoric), paleontological, historical, cultural, artistic, and religious values, as well as unique natural environmental features that embody cultural values, such as sacred groves. … intangible forms of culture, such as cultural knowledge, innovations and practices of communities B. David, B. Duncan, M. Leavesley 4-458 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project embodying traditional lifestyles, are also included. The requirements of this Performance Standard apply to cultural heritage regardless of whether or not it has been legally protected or previously disturbed. Natural features of the landscape such as specific hills, clay sources, beaches, sago stands and waterholes, which may be unmodified but nevertheless regarded as significant by local communities, are included under IFC Performance Standard 8. Such places are also of relevance in the sense that it may not be appropriate to only protect an archaeological site (e.g. a particular ancestral village site) while at the same time destroying its immediate surroundings (e.g. its associated waterhole or beach), if these immediate surroundings also form part of a site’s culturally significant matrix. 4.5.4.1 Historic and Maritime Heritage Legislation Several pieces of legislation also apply to underwater or maritime (including both pre-Europeancontact period and later) and other European-contact period sites in PNG. All the definitions of cultural property identified under the National Cultural Property (Preservation) Act listed above also and equally apply to contact period and maritime scenarios, particularly as the Act does not define the term ‘ancient’ and is not limited to terrestrial places. In particular, the specific reference to ‘a deposit of … historical remains’ (§20[1][c], which is referred to in §20[3]) indicates that all ‘historic’ and maritime sites are reportable under this Act. These sites include artillery batteries and other associated defence structures, planes and shipwrecks, graves and, similarly, all burial sites (including sites associated with World War II [WWII] conflict and crashes). The PNG National Museum and Art Gallery is empowered under §6 and §7 of the National Cultural Property (Preservation) Act to investigate any destruction or illegal export of, and to compulsorily acquire, any movable objects and immovables (the land where they are located) which have been declared to be National Cultural Property under the Act. Under the terms of the National Museum and Art Gallery Act (1992), the PNG National Museum and Art Gallery is also responsible for administering the War Surplus Material Act (1952). The War Surplus Material Act (1952) and accompanying Regulations (1952) specifically protect all material from World War II and other times of defense for the period between 1939 and 1952, including any plane or shipwreck, vehicle or machinery (such as cars, trucks, tanks and so forth) or their associated parts; unexploded ordinance; building, fitting or other structure (or its associated materials) including those situated in internal waters, territorial seas or underground. All war-related materials are deemed to be the absolute property of the State. There is a mandatory requirement (§9) that the discovery of any material of this type must be reported to the Curator of DoMH, a division of the PNG National Museum and Art Gallery. The Curator is currently Mr Mark Katakumb. The DoMH is responsible for the administration of modern periods, and specifically WWII materials and sites, and maintains a Register of military sites, relics and collections. Of note, the Act restricts any interference with unexploded ordinances (UXO) and other munitions. If UXO materials are discovered, it is a requirement that they be reported initially to the DoMH, who will organise their removal through the PNG Bomb Squad. However, UXOs can be removed by any persons discovering such a site, if the team has an approved UXO disposal expert working with them (Mark Katakumb, personal communication 2008). There are also provisions under the War Cemeteries and Graves Act (1986) for the responsible Minister to declare an area of land a Commonwealth War Cemetery. Once this clause has been enacted, it is an offence to disturb or exhume a body buried in such declared places without Ministerial approval. Many military aircraft and shipwreck sites are commonly regarded as ‘war graves’ around the world (Gibbs, 2005:60), and as such hold strong social and spiritual attachments for relatives. Such sites are often the scene of memorial services (for an apt example, see the recent press reports surrounding the discovery of HMAS Sydney). It is thus possible, and indeed likely, that newly discovered aircraft or shipwreck sites associated with loss of life could be deemed war graves and awarded suitable protection under this Act. International parties may also have an interest in such sites and consider them war graves or places of special significance (see section 4.5.4.3.1, section B. David, B. Duncan, M. Leavesley 4-459 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.4.3.2 below). It should also be noted that under the Cemeteries Act (1955), it is an offence to exhume a body from a grave, and a body still entombed in an aircraft or shipwreck might also be considered a grave. This Act is also relevant to those areas with ancient village sites containing burials, and where other kinds of burial sites (e.g. cave ossuaries) are present. The National Cultural Commission Act (1994) also contains clauses for ‘the preservation … of indigenous lifestyle of the peoples of Papua New Guinea as well as their cultural heritage [both tangible and intangible] and values’ through a cultural development programme (§1 of the Act). 4.5.4.2 Non Heritage-Specific Legislation Other PNG national legislation indirectly applies to the protection of maritime cultural heritage sites. There are a number of clauses under the Merchant Shipping Act (1975) that pertain to shipping incidents, shipwrecks and their associated fittings and cargoes. These include offences and penalties for interfering with, plundering and/or removing a wreck (§263J, §263Q); obligations to report the discovery of a shipwreck (§263K); and transferral of ownership to the State where original ownership cannot be ascertained (§263N). Under §37-43 and §47 of the National Agriculture Quarantine and Inspection Authority Act (1997), any overseas vessel or aircraft (and all their associated goods) that has not previously cleared quarantine is still subject to quarantine regulations until such time as it is cleared or given a certificate of pratique. It is an offence (under §63) to remove any material from a vessel or aircraft that has not cleared quarantine. These regulations still apply to planes and ships that were wrecked before they had cleared quarantine (particularly those that had not already docked in PNG or were flying defence sorties directly from Australia). Similarly, under the Customs Act (1951) it is an offence to import any goods that have not cleared customs through a designated port of entry. Goods that have not cleared customs within a designated period are declared forfeit to the State (§25, §25a). This Act further outlines obligations to report the incidence and/or discovery of a ship or plane wreck to a customs officer, and that it is an offence to remove, alter or interfere with any part of a wreck or its associated goods that may be subject to customs duties (§30, §31, §32). This is reinforced under §263B of the Merchant Shipping Act. Examples of these types of cultural property may be present within the project development area. This summary of PNG national legislation is indicative of what the project development area may be subject to in regard to cultural heritage protection under PNG law. 4.5.4.3 International Instruments 4.5.4.3.1 Historic and Maritime Heritage Legislation – Australian Legislation Certain PNG shipwrecks and associated relics in waters adjacent to Australian Commonwealth waters or within/above the outer limit of the continental shelf of Australia may be declared as historic shipwrecks and/or relics under the (Australian) Historic Shipwrecks Act (1976) (§5.5, §5.6). Although one of the applications of this Act appears to be to protect PNG wrecks in Australian waters, it is unclear whether or not this also applies to PNG wrecks that occur on the Australian side of the continental shelf where it is contiguous to the PNG coastline, as the Act stipulates that all wrecks that occur on the Australian continental shelf are also protected. Further work by maritime law specialists is required to determine the implications of the Petroleum (Submerged Lands) Act (1967), Indonesian Border Agreement Act (1973), Petroleum (Gulf of Papua) Agreements Act (1976), National Seas Act (1977), and Offshore Seas Proclamation (1978) in regards to any overlap between Australian and PNG sovereignty boundaries, particularly with regard to the extent of the Australian continental shelf, and the implications of these Acts on the application of the Australian Historic Shipwrecks Act (1976). The Office of Australian War Graves (OAWG) also maintains a Register of War Dead, and has an undertaking to maintain and commemorate war graves in perpetuity (OAWG, 2008). The OAWG B. David, B. Duncan, M. Leavesley 4-460 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project would show interest and concern for any graves of Australian servicemen found discovered in PNG, including those where crewmen died and were still interred in their vessels or craft. 4.5.4.3.2 Historic and Maritime Heritage Legislation – U.S. Legislation Many countries have enacted legislation to ensure that their wrecks of military craft are protected through perpetual retention of salvage rights (United States Sunken Military Craft Act §1401, §1406). For instance the United States Sunken Military Craft Act (2004) states that all sunken United States (US) military craft remains the property of the US government, and that salvage rights will never be extinguished without express divestiture of title by the United States government. In this context, sunken military craft applies to any US military ship or plane which was wrecked during military service (§1408). Furthermore, should any war aircraft crash or shipwreck site be discovered that contained human remains, it is probable that such sites would be declared war graves by their country of origin, and that protection would be sought from the host nation (in this case, PNG). This was recently the case after the discovery of the Australian Battle Cruiser HMAS Sydney and the German raider HMV Kormoran off the Western Australian coast (e.g. see Blenkin, 2008). It should be noted that the US Defence Department has a co-ordinated programme to locate, record and exhume the remains of former US servicemen for return to their families. The Joint POW/MIA Accounting Command (known as JPAC) runs worldwide operations with dedicated field teams of forensic scientists and archaeologists to achieve this aim. It has in-country detachments in Vietnam, Thailand and Laos. Japan also runs a similar program to recover the bodies of Japanese servicemen. Both of these US and Japanese agencies should be consulted wherever any allied aircraft, shipwrecks or other war grave and crash sites are discovered, and should also be notified of the aircraft wrecks identified in the present report. 4.5.4.3.3 International Conventions: Cultural Heritage as a Human Right With regard to ‘historic’ and underwater/maritime sites specifically, the UNESCO Convention for the Protection of Underwater Cultural Heritage (2001) has outlined several statements and conditions which are rapidly being accepted as global standards for the protection of submerged cultural heritage. The convention defines underwater cultural heritage as ‘all traces of human existence having cultural, historical or archaeological character which have been partially or totally underwater, periodically or continuously, for at least 100 years’, such as: • Sites, structures, buildings, artefacts and human remains, together with their archaeological and natural context. • Vessels, aircraft, other vehicles or part thereof, their cargo or contents, together with their archaeological and natural context. • Objects of prehistoric character. Pipelines and cables on the seabed are not considered as underwater cultural heritage (Article 1). The convention recommends and encourages: • In situ conservation and conformity of protection, documentation and legislation regarding underwater cultural heritage, and discourages commercial exploitation of the same (Article 2). • That any salvaging (where authorised) should conform with the protection standards of the Convention (Article 4). • The reduction of any adverse affects on underwater cultural heritage (Article 5). B. David, B. Duncan, M. Leavesley 4-461 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project • Co-operative State party and internal agreements to preserve underwater cultural heritage (Articles 6, 7 and 8). • Development of reporting procedures which ensure that the discoveries of international cultural heritage sites are made known to the international community and the site’s country of origin, to enable co-operative management/protection of those sites (Articles 9, 10, 11 and 12). • That although military and government vessels/aircraft have immunity from these conditions, they should endeavour to abide by them where possible (Article 13); • State parties should endeavour to terminate and control the illicit trade in underwater cultural heritage relics, and impose sanctions and seizure of the same where possible (Articles 14-18). • State parties should promote co-operation and information sharing through public awareness and training in underwater archaeology and through the establishment of competent authorities, cooperative meetings and mediation arrangements (Articles 19-22). • That the Convention applies equally to inland and maritime waters (Article 28). The Convention outlines rules regarding activities directed at underwater cultural heritage sites. These rules include: • In situ preservation, protection and documentation. • The restriction of sale or trade of underwater cultural heritage as commercial items. • Minimisation of the affects of activities on the underwater site. • Use of non-destructive investigative/survey techniques in preference to excavation/recovery. • Avoidance of all activities causing unnecessary disturbance to human remains or venerated sites. • Regulation of activities to ensure proper recording of cultural, historical and archaeological information. • Promotion of public access to sites. • Encouraging international co-operation to improve codes of conduct and skill-sharing. • Improved planning and research design to facilitate significance assessments, adequate funding and timetabling, expertise and onsite conservation, reporting, curation, archiving and dissemination. • Greater consideration of environmental impacts of research/investigation. Although PNG is not currently a signatory to UNESCO’s Convention for the Protection of Underwater Cultural Heritage, these standards should be adopted as best practice in any operation dealing with underwater/maritime archaeological sites. The UNESCO Convention on the Means of Prohibiting and Preventing the Illicit Import, Export, and Transfer of Ownership of Cultural Property (1970) equally applies to the trade in relics from land and underwater sites, including ship and plane wrecks. Drowned landscapes, where former terrestrial areas have been inundated, may potentially hold evidence of prehistoric occupation (e.g., as is likely to be the case shortly to the east of the Kopi Bypass corridor at the ancient village site of Areviti [site KG41]). Many previous studies (e.g. Dortch et al., 1990; Fischer, 1995) have demonstrated the diversity and remarkable state of preservation of stone and organic artefacts that might be found in inundated landscapes. While a previously terrestrial site may now be under water as a result of shifting river courses or rising sea levels, a drowned site B. David, B. Duncan, M. Leavesley 4-462 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project should be investigated to determine whether or not it still retains cultural materials worthy of future management. English Heritage, the premier heritage organisation in the United Kingdom (UK), has set out guidelines for detecting prehistoric artefacts using seabed dredging operations; such guidelines have been adopted by commercial dredging operators (British Marine Aggregate Producers Association and English Heritage, 2003). These standards include using a suitable screen to sample dredge spoil at regular intervals to determine the presence of prehistoric artefacts. 4.5.5 Methods Following the brief for this study, and in accordance with the various legal instruments (including both domestic legislation and international covenants), and ethical and professional best practice expectations outlined in Section 4.1, the cultural heritage site survey team has employed the following methodology in carrying out the cultural heritage sites impact study presented in this report. The methodologies employed in the Onshore and Offshore LNG Facilities components are presented separately in section 4.5.5.1 and section 4.5.5.2 respectively. Knowledge of cultural practices and historical trends and events for the study region is based on five kinds of sources: • Oral traditions. What living people can tell us about their past and present cultural practices, cultural events and cultural sites. • Historical records. These include early colonial (generally European) and later observations of local peoples, as recorded in written texts, historic maps and charts, photographs and audio-visual recordings (explorers’ reports, missionary accounts, government patrol reports and so forth). However, in a maritime context these also include historical documentation of past sailing routes and port facilities (such as pilots’ records and/or sailing directions; notices to mariners; governmental public works records; Customs Department records of shipping incidents). WWII sites are documented through numerous avenues, including defence department and other administrative records of WWII air and shipping losses; civilian registers of aircraft crashes; regional maps and plans. • Anthropological and linguistic writings. What anthropologists and linguistics have to say about cultural heritage matters and cultural sites, based on professional participant observations and linguistic analyses. • Museum and other portable objects. Artefacts which came from particular locations on the land or in water and which help give those places distinctive cultural significance. • Archaeological evidence. Material remains (artefacts) of past human activity obtained through archaeological investigation. Throughout this report, unless otherwise indicated all locations are given in the WGS84 (Zone 55) coordinate system. 4.5.5.1 Onshore LNG Facilities Component The archaeological team undertaking the LNG152 Land component fieldwork consisted of Dr Bruno David – Project Co-ordinator, archaeologist (Monash University, Australia), Dr Matthew Leavesley – Field Survey Co-ordinator, archaeologist (University of PNG, Port Moresby), Nick Araho – Archaeologist (PNG National Museum and Art Gallery, Port Moresby), Jeremy Ash – Archaeologist (Monash University, Australia), John Dop – Cultural heritage officer (PNG National Museum and Art Gallery, Port Moresby), Dr Brad Duncan – Archaeologist (Monash University, Australia), Dr Alexandra Gartrell – Cultural geographer (Monash University, Australia), Alois Kuaso – Archaeologist (PNG National Museum and Art Gallery, Port Moresby), Herman Mandui – Archaeologist (PNG National Museum and Art Gallery, Port Moresby), Moi Dobi – Community representative (Boera village), Gau B. David, B. Duncan, M. Leavesley 4-463 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Ario – Community representative (Papa village), Renagi Koiari – Community representative (Papa village). th th The fieldwork for this report began on 17 January and ended on 11 May 2008. It was undertaken th st during two phases, the first from 17 January to 21 February when surveys began at the northwest 2 st th corner of a 8.18 km kidney-shaped area within Portion 152; and the second from 21 April to 11 May 2 of a smaller, 2.64 km area (from here-on the ‘Survey Focus Area’) largely but not exactly contained within the original kidney-shaped area. As noted in section 4.5.3.1, the reason for these two phases is th that an UXO was found by the field crew during the first period of fieldwork, on 20 January. Upon st reporting this discovery to Coffey Natural Systems on 21 January, the cultural heritage site survey team was instructed for safety reasons to cease all Onshore LNG Facilities component fieldwork until further notice. This arrived during a series of communications from Coffey Natural Systems between nd th th 2 and 30 April (especially 25 April), when the brief was changed and the Monash University cultural heritage team was requested to survey only along 3 m-wide transect lines pre-cleared by MilSearch personnel within the area of the Survey Focus Area (because the rest of the Survey Focus Area had not been cleared of UXO by MilSearch in time for 100% archaeological surveys to proceed). 2 2 As a result, the western 39% (0.89 km ) of the total 2.28 km of the Survey Focus Area inside the site security fence area was entirely (100% ground coverage) field surveyed during the first phase of fieldwork, but the rest of the Survey Focus Area was only sampled along the transect corridors (Figure 4.5.5). Figure 4.5.5 Areas surveyed in January (100% surveys) and April-May 2008 (sample transects) within and near the site perimieter fence and Survey Focus Area. 2 Additionally, an area of 0.75 km to the immediate northwest of the Survey Focus Area was entirely (100%) surveyed, because the original kidney-shaped area prior to the discovery of the UXO included an area extending outside the subsequent Survey Focus Area. In June, Nick Araho undertook an extra day of fieldwork to record two ancestral village sites outside th the development area, Aemakara and Dirora. On 19 June, five days before this report was due for B. David, B. Duncan, M. Leavesley 4-464 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project submission, the cultural heritage team was given a map showing the final proposed development footprint (the area inside the ‘site security fence area’, being the security fence at the outermost boundary of the LNG Plant site) by Coffey Natural Systems. The area inside the site security fence 2 area covers 7.93 km , 29% of which is covered by the Survey Focus Area (Figure 4.5.2). In this report (see section 4.5.11 in particular), cultural heritage site distributions are presented for the area inside the site security fence area, based on field surveys undertaken in the Study Area. Impacts of the proposed developments on cultural heritage sites within the site security fence area are presented in section 4.5.13.1, and recommendations in section 4.5.14.2.1. Interviews and site visits were undertaken with local Koita and Motu community members at a series th of public meetings held at Porebada, Boera, Papa and Lea Lea villages on 25 January 2008, and th th each day from 6 to 10 May 2008. Additionally, field participants nominated by village representative bodies participated in the archaeological fieldwork during both the January and April-May surveys (see the above bullet-point list for the names of each representative; see section 4.5.4.1 for specific details). The following people participated at these community discussions in public meeting places and/or as individual discussions at Porebada, Boera, Papa and Lea Lea: Aia Avata, Daro Avei, Rakatani Henri, Diari, Siosa, Lohia Gabe, Tara Gau, Mea Gudia, Billy Heni, Iru Kari, Robert Kauga, Maba Lohia, Oveai Maino, Igo Meauri, Lohia Miria, Mea Miria, Nadani Morea, Reverend Vagi Naime, Heni Totona, plus another approximately 30 people who attended these meetings but whose names were not recorded. During the second phase of fieldwork, Jessica Wiltshire and Robert Bone of Coffey Natural Systems joined the Monash University cultural heritage team as observers. 4.5.5.1.1 Historical Sources Investigated A number of documentary sources were investigated for background information on cultural heritage and cultural heritage sites in and surrounding the Study Area. These include the site register and reports files of the PNG National Museum and Art Gallery, a limited investigation of cultural items held by the PNG National Museum and Art Gallery, and a detailed review of the published literature on the archaeology of the southern PNG lowlands, focusing on the Port Moresby area. 4.5.5.1.2 Fieldwork and Site Recording Methods All cultural heritage sites identified during the fieldwork within the Study Area were recorded on recording forms specifically designed for this cultural heritage consultancy (the original recording forms were slightly modified after the January fieldwork and used in the April-May fieldwork; see Annexure 4.5.1). All sites were recorded in consultation with representatives from Papa and/or Boera villages. Cultural heritage sites were deemed any place showing evidence of cultural heritage, be it a large village, a sacred place or an isolated stone artefact. Evidence for the presence of cultural heritage sites followed two principles: • Archaeological sites were by definition identified through the presence of material remains relating to past cultural activities (e.g. stone artefacts, pottery sherds or shells). Neighbouring archaeological sites were differentiated when artefact concentrations were separated by areas devoid of surface culture materials more than 5 m apart. • Oral tradition sites are any cultural heritage site known from oral traditions. In principle, an oral tradition site could be a sacred site lacking material evidence of past human presence (i.e. a cultural heritage site that is not also an archaeological site), or it could also be an archaeological site (i.e. it may have an anthropogenic material signature), but in practice we found that all of the oral tradition cultural heritage sites communicated to the cultural heritage site survey team by community representatives were outside the Survey Focus Area. That is, all cultural heritage sites within the Survey Focus Area are archaeological sites. B. David, B. Duncan, M. Leavesley 4-465 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project The survey was undertaken in two stages. Due to the complexities resulting from the discovery of the UXO during the January survey, the methodologies for each survey were markedly different and therefore are described separately below. th th January Survey. The survey was undertaken on 17 to 20 January 2008. The participants for this survey included Matthew Leavesley, Herman Mandui, Nick Araho, Alois Kuaso and Moi Dobi in Team 1; and Brad Duncan, Jeremy Ash, Jon Dop and Gau Ario in Team 2. The January survey strategy was designed to identify every archaeological object on the landscape in the Priority Area including isolated (individual) midden shells and stone artefacts. In January 2008, prior to the discovery of the first UXO, we undertook archaeological field surveys in the following way: 1. Prior to commencement of fieldwork, a meeting was convened with Herman Mandui (Chief Archaeologist, PNG National Museum and Art Gallery), Nick Araho and Alois Kuaso of the Prehistory Department at the PNG National Museum and Art Gallery (which administers the National Cultural Property [Preservation] Act). Alois Kuaso undertook to gather documents from all known archaeological surveys previously undertaken within the study region for photocopying. The locations of existing archaeological sites were plotted on 1:50,000 topographic maps, and individual site records were photocopied from the Museum files. th 2. On the first day (Thursday 17 January), with the assistance of Esso Highlands Ltd. Community Affairs officer Agi Hoire, we travelled to Boera and Papa to liaise with the community leaders so as to meet with the community representatives from the respective villages to assist with the cultural heritage surveys. We then identified the geographical parameters of Portion 152 and the proposed archaeological survey area within it. This included a cursory driving tour of parts of Portion 152 and a visit to Konekaru beach. The information obtained allowed us to formulate an efficient survey plan th th th for the remainder of the survey. On the 18 , 19 and 20 January we divided into two teams (Matthew Leavesley, Herman Mandui, Nick Araho, Alois Kuaso and Moi Dobi in Team 1; Brad Duncan, Jeremy Ash, John Dop and Gau Ario in Team 2). th On Friday 18 January we operated as two co-ordinated teams. Team 1 began an intensive survey while Team 2 recorded the large site complex (and associated nearby sites) identified the previous day. Team 1 began on the junction of the northern boundary of Portion 152 and the Boera-Papa Road. The team formed a line with each individual spaced 10-15 m apart and simultaneously walked west to the coast. When a site was identified the team stopped to identify its respective parameters and record the site. Upon completion of the recording process the line was reformed and the team continued to walk across the landscape. Once the team reached the western boundary of Portion 152, it reformed south of the southern-most surveyor and return back (east) across the landscape to the road. This procedure altered slightly at the junction of the Boera-Papa Road and Ruisasi Creek. Rather than crossing the creek with every sweep, the team remained on its western side. This process was continued for the remainder of the survey. th From the outset, Team 2 followed a slightly different strategy. It began Friday 18 by returning directly to the large site complex at Konekaru beach and recorded all of the sites within the vicinity. Upon completion of this task the team moved to the dry ground north of the confluence of Ruisasi Creek and Vahui River. Team members recorded all the known sites in the locality and then formed their own line of surveyors spaced 10-15 m apart and began to survey on an east-west axis moving to the north, back towards Team 1. 3. Although ground visibility was in the main between 0-25 % due to long grass on the field and thick mangrove cover on the coastal margin, conventional ground-walking archaeological surveys were employed for this study. Significantly, the one old village location identified as Konekaru and lying along the beach where the pipeline was at one stage proposed to come onshore was identified by B. David, B. Duncan, M. Leavesley 4-466 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project village representatives during our field surveys, without us ever having to ask local people about the locations of individual villages – that is, the village was memorialised in local clan oral histories, and was communicated to us during the ground surveys. The villagers who showed us cultural heritage sites during this phase of the project are the following: Moi Dobi (from Boera) and Gau Ario (from Papa). 4. While travelling to and from cultural heritage sites shown to us by clan representatives, walking tracks, creek banks, mudflats, uprooted tree hollows, and the sediment matrix trapped in tree roots were examined for archaeological remains, and all sites thus found were recorded. 5. Whenever we came across a site, whether taken there by local community representatives, or by being told of nearby sites presently inaccessible due to thick vegetation cover, or by finding isolated stone artefacts along tracks/tree roots/creek banks while walking to other sites, an individual Site Recording Form was completed (4.5 Annexure). GPS locations were recorded for each site we went to, and estimated using compass bearings for more distant sites we could not directly access; locations at these latter sites were identified from the 1:50,000 topographic map following informant interviews conducted in proximal locales. 6. At the completion of the January survey, the team had all but completed an intensive survey of the region bounded on the north by the boundary of Portion 152, the east by the Papa-Boera Road and Ruisasi Creek, and the south and west by the mangroves. 2 2 Using these methods, a total area of 1.64 km covering the western 11% (0.89 km ) of the Survey 2 Focus Area plus an area of 0.75 km immediately outside but contiguous with the Survey Focus Area (to the northwest) were 100% surveyed in January 2008. The total contiguous area systematically 100% surveyed within and adjacent to the site security fence area is equivalent in size to 21% of the site security fence area. st th April-May Survey. This survey was undertaken between 21 April and 8 May 2008, and was conducted very differently to the January survey. Under the directions of Coffey Natural Systems and Esso, the survey was restricted to the areas previously cleared of UXOs by MilSearch. Site identifications proceeded in the following way: st 1. Prior to commencement of fieldwork on Monday 21 April, we liaised with Esso Highlands in order to participate in a site induction conducted primarily by MilSearch. 2. Upon arrival at Portion 152 we liaised with MilSearch and the local village representatives and immediately divided into two surveying teams and remained so for the duration of the survey. Team 1 consisted of Matthew Leavesley, Herman Mandui, Nick Araho, Jessica Wiltshire (Coffey Natural Systems) and Moi Dobi. Robert Bone (Coffey Natural Systems) also briefly joined Team 1. Team 2 consisted of Jeremy Ash, John Dop and Renagi Koiari. Each team was also accompanied by a MilSearch representative. 3. The survey was limited to narrow transects and areas of high ground visibility that had previously been cleared by MilSearch. This included the transects, mudflats and other areas clear of long grass. In the latter case, the MilSearch representatives cleared the area before we ventured into it. All but one small section of the transects were 3 m wide. The only exception to the use of the transects occurred when either a MillSearch representative cleared a path in front of us, or on open ground (such as the mudflats) where the MilSearch representative could clearly see the ground ahead of us. The spatial distribution of the transects was designed to ensure that all of the major topographic features on the landscape were encountered at least once. They were also spaced in such a way that large archaeological sites would not be missed. The transects were divided between the two teams and were intensively surveyed. These factors seriously limited the scope of the survey (i.e., the eastern part of the Priority Area has not been 100 % surveyed). B. David, B. Duncan, M. Leavesley 4-467 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4. Although ground visibility was low, conventional ground-walking archaeological survey was employed for this study. The two teams intensively surveyed the ground, within the requisite areas delineated by MilSearch, for archaeological material. Whenever we came across a site, an individual Site Recording Form was completed (4.5 Annexure). GPS locations were recorded for each site we went to, and estimated using compass bearings for more distant sites we could not directly access; locations at these latter sites were identified from the 1:50,000 topographic map following informant interviews conducted in proximal locales. The cultural heritage site recordings were not aimed at mapping present-day land-use patterns, but historical sites that may potentially take a number of culturally meaningful expressions in the landscape. This might include the Department of Primary Industry (DPI) farm that was set up early last century (post-1906). Thus, the kinds of sites recorded are those sites that allow people to trace the present with the past. Such sites may include conventionally defined archaeological sites – that is, sites that have physical and potentially datable evidence of past activities, such as stone tools, pottery or wooden structural remains – or they may be ancestral or ‘origin’ sites for local village members or sacred sites which allow local people to spiritually connect with ancestral homelands (as per oral traditions). Thus present-day gardens, villages or bush camps have not been mapped in this cultural heritage study unless these are also said to have been historically significant places by clan members, or showed evidence of archaeological significance during the course of the surveys. Public and individual discussions and interviews were also held at Porebada, Boera, Papa and Lea Lea villages to ensure that cultural heritage sites within the Study Area known by local villagers were documented. During these interviews, various issues and concerns were raised by Koita and Motu individuals and representatives. Additionally, village representatives took members of the cultural heritage team to cultural places outside, but close to, the Survey Focus Area for field recording. These sites and issues are documented in various parts of this report, in particular section 4.5.10 and section 4.5.14 (survey results and recommendations, respectively). 4.5.5.2 Offshore LNG Facilities Component It is important to note that this is the first time that formal maritime archaeological research has been undertaken in mainland PNG for EIS purposes. Therefore, there are no previous compilations that could be used as foundation documents, as this study entailed locating and exploring information sources previously not considered or consulted. The location, extent, availability and accessibility of these data sources varied greatly. In this way, the shallow marine investigations carried out during this consultancy set a benchmark for future consultancies. The archaeological team undertaking the Offshore LNG Facilities component fieldwork consisted of Dr Brad Duncan – Co-ordinating archaeologist and commercial diver (Monash University, Australia), Lyall Mills – Commercial diver, commercial dive supervisor and avocational archaeologist (Monash University, Australia), Scott Allen – Remote sensing specialist, commercial diver and avocational archaeologist (Monash University, Australia), Liz Kilpatrick – Archaeologist and commercial diver (Monash University, Australia), Gau Ario – Community representative (Papa village), Auda Delena – Community representative (Lea Lea village), and Moi Dobi – Community representative (Boera village) 4.5.5.2.1 Historical Sources Investigated Given an absence of pre-existing synthesized information or bibliographies for the Study Area, several local knowledgeable institutions were approached by which to establish a background history for the Caution Bay area. The archival collections of the following institutions were investigated for this purpose: • PNG National Museum and Art Gallery (Department of Prehistory), Port Moresby. • PNG National Museum and Art Gallery (Department of Modern History), Port Moresby. B. David, B. Duncan, M. Leavesley 4-468 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project • History Department of the University of PNG, Port Moresby. • Archaeology Department of the University of PNG, Port Moresby. • Ports PNG, Port Moresby. • PNG National Archives, Port Moresby. • National Research Institute, Port Moresby. PNG National Museum and Art Gallery (Department of Prehistory). The Prehistory Archaeological Database was searched with the assistance of Curators Herman Mandui and Nick Araho. There were no known records of indigenous archaeological sites in the underwater section of the Study Area. PNG National Museum and Art Gallery (Department of Modern History). The Department of Modern History (DoMH) maintains a database of aircraft and other military wrecks and relics from WWII. DoMH also houses an extensive collection of military and civic maps, plans and charts, books, and databases of missing WWII aircraft. The entire map collection, missing aircraft database, and selected archival documents and publications were investigated with the assistance of the Curator Mark Katakumbe and Technician John Lelai. Archival research was concentrated in this facility to begin with, as the most likely anticipated maritime sites in the Study Area were historical shipwrecks, aircraft wrecks, defence sites and other historic infrastructure. Archaeology and History Departments, University of PNG. Interviews were conducted with Dr Matt Levingsley (Archaeology) and Professor Biama Kanasa (History) to ascertain the extent of previous research in the Caution Bay area. Biama Kanasa has undertaken historical research in this area (with particular reference to oral history documentation), and during the course of the archival research for the present Consultancy discussed his findings and previous publications with Brad Duncan. However, no previous maritime or post-European contact period archaeological investigation had been undertaken in the region. Ports PNG. Ports PNG control the movement of shipping in and out of all ports, and provide sea pilots to navigate large vessels. Although the organisation did not retain an archive of Notices to Mariners (as was anticipated), local sea pilots did have a detailed knowledge of the coastal waters and provided Brad Duncan with information not otherwise available, particularly regarding maritime defence facilities and the absence of shipwreck sites in the Caution Bay area. National Archives/National Library. Unfortunately, the National Archives and the PNG National Library were both closed during the period allocated for historical research. The latter was closed indefinitely for major refurbishment. Although one day’s research of the air crashes and airstrips folders (SN54) was subsequently undertaken on our behalf by a local contractor when the National Archives re-opened, results were negligible. Further research in the archives may reveal historical data pertinent to the Study Area. Areas recommended for further research include the following Series: • Aerodromes and Air Services (GSS360) – includes air crashes and airstrips. • Air Services (SN19). • Airfields and Airline Services (SN54) – includes reports of air crashes. • Allied Geographical Studies (SN94) – includes aerial photography of military forces activities in WWII. • Armed Services (SN55). • Bomb Disposal (SN55). B. David, B. Duncan, M. Leavesley 4-469 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project • Cemeteries (SN45) – includes Commonwealth war graves and Japanese graves. • Cemeteries (GSS365) – includes graves of National and Australian servicemen; Customs and Trade (GSS370). • Defence (GSS400). • District and Departmental Annual Reports (SN100). • Folding Maps - British New Guinea (SN664). • Lands, Surveys and Mines (SN70). • Patrol Reports (SN659) – includes social, economic and political reports 1900-1980. • Public Services (GSS395) – includes wharves and jetties. • Scrap Metal Disposal (SN24). National Research Institute. The National Research Institute holds an extensive collection of PNG literature and historical records in their library. A brief search of this collection for primary shipping records and WWII defence records did not reveal any relevant sources, although more detailed research in the institution’s library may hold hitherto unknown details. PNG Department of Defence. Planned research with the PNG Department of Defence did not take place, due to difficulties in arranging meetings with senior Defence Department personnel. The PNG Defence Forces may hold an archive including information applicable to this study. However, given the Australian administration of PNG until 1975, it is likely that most documentation relating to military activities during WWII is held at the Australian War Memorial Archives (see below). Other PNG Sources. Other potential PNG avenues of enquiry identified but not explored due to time constraints included divers and seamen associated with The Underwater Explorers Club of PNG (Port Moresby Branch), and the Port Moresby Yacht Club. Australian War Memorial. The Australian War Memorial (Canberra, Australia) holds extensive records regarding Australian and Allied wartime activities. Selected pictorial collections have been searched on-line (see http://www.awm.gov.au/database/cas.asp), although further investigation of these sources is recommended. Details of files relevant to PNG military activities are outlined in Kanasa (1996). National Archives of Australia. Previous research by Mr Peter Taylor of the Maritime Archaeological Association of Victoria (Australia) of the National Archives of Australia (Victorian Branch) had identified files regarding WWII war wrecks in PNG. However, no information regarding ship or plane wrecks in the Caution Bay area was contained in these files. 4.5.5.2.2 Fieldwork and Site Recording Methods The cultural heritage sites documented in this report occur within the area known as Caution Bay, which extends from Redscar Head in the north, to Idihi Island in the south and all the areas east to Boera Head. Sites were deemed to be any place showing physical evidence of cultural heritage, regardless of size or antiquity. Site identification proceeded in the following way: Oral Histories. Oral histories form a significant resource for identifying archaeological and traditional sites, and can often be utilised to definitively locate previously undocumented maritime sites (see Duncan, 2007:44). Prior to the commencement of fieldwork, oral histories were collected from the three villages located close to the Study Area. Starting from the north proceeding southwards, these villages were Lea Lea, Papa and Boera. Interviews were undertaken with fishermen and other local people who had a demonstrated knowledge of the offshore regions. Dr James Weiner (personal B. David, B. Duncan, M. Leavesley 4-470 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project communication 2008), who had been undertaking a concurrent social anthropological survey of these villages, had observed that the people of Boera and Lea Lea made greater economic use of the ocean than those at Papa do, as the former villages had predominantly fishing-based economies, whereas the latter was more terrestrially based. Hence, the collecting of fishermen’s oral histories for purposes of this report was concentrated at Boera and Lea Lea, with a greater emphasis on terrestrial areas and mangroves undertaken at Papa. Fishermen in all three villages demonstrated in-depth knowledge of the underwater landscape in the areas utilised for their fishing activities. Most of the fishing in this region is today undertaken by freediving from small outrigger canoes to spear fish and crayfish, and to collect molluscs (giant clams and sea urchins) and bêche-de-mer (sea cucumbers). Consequently, local fishermen possess an intimate knowledge of the submerged topography and reefs, along with archaeological sites found in those locations. It should be noted that as underwater wreck sites frequently provide an attractive habitat for fish, it was highly probable that local fishermen would know the location of any wrecks in the region. Three community representatives (one from each village: Auda Delena at Lea Lea, Gau Ario at Papa and Moi Dobi at Boera) accompanied the fieldwork crew on a boat-visit to the Study Area to locate archaeological sites. Later, after consultation with village elders and other fishermen/villagers, the representatives produced a map of Caution Bay showing the location of known archaeological and traditional sites, along with the names of local reefs used and associated with traditional fishing rights (Figure 4.5.6). Prior to the commencement of fieldwork, a number of researchers and government officers who had detailed knowledge of the Study Area and now resident in Port Moresby were interviewed to collect oral histories of any activities they had undertaken in and around the Study Area, as well as their knowledge of potential sites. These people included: • Environmental consultant (John Douglas). • Local avocational archaeologists and divers (Neil Whiting, John Miller). • World War II veteran from Papa (Ben Moidé). • Members of the PNG bomb squad (Lt Tui Gaileko, Petty Officer Steven Yamun, Leading Seaman Ausa Tau and Sub-lieutenant Allan Mitmit). • Ports PNG pilot (Captain Charles Kabilu). B. David, B. Duncan, M. Leavesley 4-471 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.6 Map of traditional places and known archaeological sites produced by representatives of Boera, Papa and Lea Lea. From map drawn by Auda Delena, Gau Ario and Moi Dobi. Red areas are locations of locally known ship or plane wrecks. All informants were asked to indicate the position of sites that they knew of on copies of Aus Chart 379. Informants were also requested (where possible) to accompany the survey team to locate and document sites, and GPS co-ordinates were taken for such sites. Where the site has not been visited in this current study, or the location could not be confirmed by survey, the approximate location as indicated by the informant has been provided. Remote Sensing Fieldwork: Magnetometer. The fieldwork used two kinds of remote sensing equipment, magnetometre and Side-scan Sonar. Here we give some basic details of how these instruments work, for it is the first time that they have been used in EIS consultancies in PNG, and understanding their potentials and limitations is important for the interpretation of results. A proton magnetometer is used to detect magnetic anomalies on or immediately below the seabed. A proton precession magnetometer works through its utilisation of the precession of spinning protons or nuclei of hydrogen atoms within a sample of hydrocarbon fluid to measure total magnetic intensity. The spinning protons are contained inside a sealed unit (usually referred to as the ‘fish’) towed behind B. David, B. Duncan, M. Leavesley 4-472 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project a vessel. These spinning protons inside the fish behave as small magnetic dipoles, which are temporarily aligned to a uniform magnetic field generated through the application of a current through a coil inside the unit. When the current is removed, the spin of the protons causes them to precess (i.e. slowly spin) about the direction of the ambient or earth’s magnetic field, in the same way that a spinning top moves around the earth’s gravity field. A small signal is then generated by the precessing protons in the coil, proportional to the total magnetic field intensity for the area at that moment (Breiner, n.d.:3). A simplified explanation of this process is as follows: when the induced magnetism from the current is switched off, the free floating protons inside the ‘fish’ align themselves in the direction of the magnetic polar alignment of any anomalies encountered beneath it. As the unit passes over an anomaly, there will be a sharp increase in magnetic readings in one direction as it passes over the magnetic extremity (or pole) of the object. The direction will return to the centre as the ‘fish’ passes the object’s middle, and then move to another sharp peak in the opposite direction as it passes the corresponding magnetic pole. These fluctuations give indications of the strength of magnetic anomalies, which are proportional to the distance from, size and/or metal content of the object encountered (Figure 4.5.7). Figure 4.5.8 is a nomogram used to determine the size of search lanes based on the possible target size. Note that 1 pound of iron 10 m away from the magnetometer will give the same reading as 1 ton of iron that is 38 m away. Therefore, although a magnetometer can detect the presence of an anomaly, it cannot independently determine the size of the object or its distance away from the unit. Consequently it is imperative to record the depth of the area being investigated in order to be able to interpret the output data. B. David, B. Duncan, M. Leavesley 4-473 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.7 Demonstration of magnetometer readout as the unit passes over a site. Note: high readings as the unit passes over each of the object’s magnetic polar regions. Figure 4.5.8 Nomogram showing the relationship between signal strength (nano teslas) and distance from the magnetometer head (from Breiner, no date:43). B. David, B. Duncan, M. Leavesley 4-474 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Magnetometers can detect magnetism and anomalies not associated with cultural heritage sites, such as those created by the surface and subsurface geology (especially igneous formations), as well as by outliers or sediments derived from these. Because the magnetometer reads anomalies relative to the background magnetic field intensity, the nature of the geology and geomorphology of a Study Area may reduce or effectively mask any anomalies resulting from cultural features. It is thus important to interpret signals in relation to those of the background geology. During fieldwork, a Wreckhunter Mk5 proton magnetometer was towed approximately 50 m behind the boat. This unit has a range of up to 100 m on either side of the magnetometer head dependent on distance from the target (usually related to water depth). A paper-trace readout recorded the geographical extent of each anomaly and was continually monitored in real time during the survey in conjunction with the side-scan sonar survey results (see below). Remote Sensing Fieldwork: Side-scan Sonar. Side-scan sonars will detect any structures and objects which protrude above the seabed. The side-scan unit works by projecting a sonic beam from a transducer at an angle towards the seabed. When an object protruding above the seabed is encountered, the beam reflects back to the fish head (see Figure 4.5.9). However, subsequent sonic beam pulses (from the ‘fish’) will continue beyond the raised object until they hit the seabed, where a similar return signal will be reflected back to the ‘fish’ head. Analysis of the returned beam signals produces a sharp solid reading for the raised object itself (usually a bright white readout) with a dark ‘shadow’ behind the object. The height of the anomaly can be calculated by an equation which measures the distance from the unit to the object proportional to the distance of the shadow (see Fish and Carr, 1990:84). Figure 4.5.9 Demonstration of side-scan sonar principles of use (from Fish and Carr, 1990:84). The simplest way to visualise this concept is to compare the side-scan to the effects of shining a torch on an object. The front of the object will be brightly illuminated by the beam, while a dark shadow whose length will correspond to the distance away from and angle of the beam and the size of the object will be produced behind it. Examples of the expected types of sites that might be found in the Study Area (and which were recorded elsewhere by us using the same equipment used in this fieldwork) are shown below (Figure 4.5.10). The side-scan unit was tested in Melbourne prior to fieldwork to calibrate the unit signals against known underwater site types. These include aircraft crash sites, along with intact shipwrecks and wreckage. As can be seen, these are very distinctive signatures which can be easily distinguished against other potential features such as reefs. B. David, B. Duncan, M. Leavesley 4-475 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.10 Left: Aeroplane wreck, Port Phillip Bay, Victoria (Australia). Note the shadow in the middle of the white area (fuselage) caused by an upright wing stump. Middle: Ship wreckage site, Port Phillip Bay. Right: Intact shipwreck in Yarra River, Melbourne. In the Offshore LNG Facilities component of the field surveys, a Burton M1 side-scan sonar ‘fish’ (Plate 4.5.1, left) was trailed approximately 15 m behind the boat at a depth of 1-2 m below the surface. This model of side-scan only reads to the port side of the ‘fish’ head (single beam only) and, hence, lane spacing was adjusted accordingly to ensure complete coverage of each area. The equipment was set to read up to 65 m from the ‘fish’ head, which affords a seabed coverage of up to 63.25 m in 13 m-deep water. Side-scan sonar readings were linked to an internal GPS unit to determine the exact location of the ‘fish’ head. Screen readout was constantly monitored during survey runs to detect any anomalies (Plate 4.5.1, right). Later interpretation of these results were undertaken using Lowrance side-scan sonar (version 1.2.2) interpretation software, which provided clearer images of anomalies than the Burton M1 unit monitor. These images were used to determine the location of each anomaly from the output data. The location of each anomaly was calculated by extracting the ‘fish’ location from the sonar readout, plotting the depth from GIS chart coverage, and then calculating the position of the anomaly on the seabed based on depth and diagonal distance from the sonar. The results were plotted using GIS software and subsequently physically investigated by divers. Plate 4.5.1 Left: Side-scan sonar ‘fish’ rigged for shallow water in the dinghy. Right: Sidescan sonar monitor. Limitations of Remote Sensing Surveys. As noted in section 4.5.6.7 below, in highly dynamic areas with a hard substrate (in this case coral and reef), archaeological sites such as planes and shipwrecks are likely to be flattened and scattered over the seabed (Riley, 1997). This observation has implications for detection of sites using side-scan sonar, as archaeological sites on high-energy reeftop areas may not be distinguishable from the surrounding coral structure. In tropical areas where coral growth is often accelerated, archaeological sites can be completely subsumed by coral in relatively short timeframes, often masking their presence completely. B. David, B. Duncan, M. Leavesley 4-476 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Although archaeological sites may survive relatively intact in cases of a softer seabed matrix (such as sand, silt or mud), they may not be visible or detectable above the seabed if they are covered completely. Furthermore, long-shore drift, currents and extreme weather events (storms and the like) can also periodically strip and/or cover large areas of seabed. Brad Duncan has observed numerous cases where shipwreck and maritime infrastructure sites have been suddenly revealed after large storms, only to be buried again within the space of a few days (e.g. the Port Albert Unidentified Wreck, documented in Duncan, 1995). In these cases, although sites may not be evident with a side-scan sonar, it may still be possible to detect associated magnetic items (if any exist) which lie either on or below the seabed surface. However, it should be noted that if there are no magnetic anomalies associated with a buried archaeological site (e.g. a timber shipwreck with no iron fittings onboard), it may prove impossible to detect it without other more localised and intensive remote sensing techniques (e.g. metal detector/sub-bottom profiling), or invasive techniques such as excavation. To further complicate the situation, even if anomalies are detected via remote sensing, water turbidity and low visibility may still affect the ability of divers to adequately locate, identify or record associated archaeological sites. It should therefore be noted that even though an area may have been surveyed using side-scan sonar, magnetometer or visual inspections, the possibility always exists that undetected archaeological sites may exist below the seabed. It is usually not possible to locate low-density pottery scatters with a magnetometer or side-scan sonar as they do not exhibit high profiles (visible to side-scan sonar) and do not produce magnetic signatures (detectable by a magnetometer). However, high-density pottery scatters can be detected by magnetometer surveys, as was recently the case near Epemeavo village in the Gulf Province (Ian Moffat, personal communication 2008). Remote Sensing Surveys: Survey Lanes Methodology. In deeper water (between 5 -13 m) the sidescan sonar and magnetometer units were towed behind the mother-ship (MV Lauta Marata), with enough spacing between them to ensure that the side-scan unit was not influencing the magnetometer readings. The side-scan was trailed 10 m with the magnetometer at 50 m behind the vessel. When in shallower water (<5 m), the side-scan sonar was towed on an extension pole on the port side of a 6 m runabout approximately 0.5-1.5 m below the surface, with the magnetometer trailed at the rear of the vessel. Search lanes were spaced every 100 m in a north-south direction and plotted using a GPS system. Transit marks were also established onshore at the extremities of the search area as a visual verification of the survey lane start and finish marks. The small vessel then towed the equipment along the survey grid lines from north to south, turning 180° and returning on the same line from south to north. This method ensured that the survey coverage included both sides of the lane (to ensure complete side-scan sonar investigation) which effectively resulted in 50 m spacings between runs, with approximately 15 m of overlap between each survey run. For the westernmost lane of each survey area, only the inshore side of that lane was surveyed (e.g. on the western extremity lane for the MOF, only the inshore side was surveyed as it was at the extremity of that buffer area). B. David, B. Duncan, M. Leavesley 4-477 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.11 Left: Remote sensing equipment set-up in dinghy. Side-scan sonar (bottom), and magnetometer on console. Right: Areas covered by maritime remote sensing survey lane runs (shown on Aus Chart 379). All of the areas in the 400 m buffer option of both the jetty and MOF sites were surveyed using remote sensing equipment. Part of the 700 m buffer section of the jetty site (mainly to the north and south of the 400 m jetty buffer, on the western side those areas) was also surveyed using remote sensing equipment (see Figure 4.5.11). Inshore areas of the 700 m option were not surveyed close to shore due to lack of time, severely restricted favourable weather conditions, and extremely low spring tides. The side-scan survey sonar and magnetometer runs were started/ended simultaneously to enable correlational comparison of data from each remote sensing source. Data could then be compared both when monitoring output during the survey in the field, and for any anomalies identified during postprocessing of the survey results. Generally, if a large anomaly was encountered with the magnetometer, it could be checked instantaneously with the side-scan sonar to verify if an extant object was visible on the seabed, and vice versa. Sea swell, wind and underwater obstructions, along with the ability of the boat driver to steer a straight course, all affected the accuracy of the lane runs. Where rough weather affected the ability of the boat driver to steer a straight course, the survey was moved to the inshore areas inside the fringing reef where water conditions were usually calmer. If it was too rough to steer a straight course in these conditions, then the survey was abandoned. In these scenarios, diving was undertaken on anomalies if safe to do so, although the adverse weather usually also prevented this. Where survey runs were interrupted by underwater obstacles, the boat was navigated around the impediment, which altered the survey run lane by a maximum of 3 m to one side. Where peninsulas (such as at the MOF site) divided the survey lane run, these survey runs were terminated when water depths became too shallow for the boat to safely operate on either side, and visual inspections of these areas were undertaken using divers on snorkel. Remote Sensing Surveys: Constraints. Many problems were experienced due to the timing of the morning surveys, which usually fell at extremely low spring tides (which hampered survey access to the inner reef due to insufficient water depth for the boat), and the occurrence of strong mid-morning to afternoon northwesterly sea breezes (of up to 40 knots) which produced unfavourable (and often unsafe) survey conditions. This severely restricted the time available to undertake the survey. Mechanical failures in outboard equipment supplied for the fieldwork further reduced available working hours. B. David, B. Duncan, M. Leavesley 4-478 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Furthermore, the time taken by the slow larger work vessel (Lauta Maurata) to access the site each morning also restricted available working hours (daily travel time was up to 5 hours dependent on weather). It is recommended that any future survey works in this area should be timed to coincide with more favourable weather conditions (possibly in the southeasterly season), and that a faster working vessel be accessed to increase productivity for working hours available onsite. Due to these limitations, underwater inspections of anomalies were restricted to the 400 m Buffer Zone for the deep water areas (over 9 m) especially as weather restrictions limited available diving time. Remote Sensing Surveys: Underwater Inspections. Wherever anomalies were detected using the remote sensing equipment, their positions were plotted with GPS and then inspected by professional (commercial) divers with archaeological experience. As uncorrected GPS has an accuracy of ±15 m, divers carried out a circular swim-line search around the derived location for each anomaly (using a 30 m tape measure attached to a weight to measure each circular line swim-line spacing) until the anomaly was either located or the extremity of the tape was reached. In most cases, anomalies were encountered in shallow waters (usually coinciding with shallow reefs) and were investigated by divers using snorkel. Where weather conditions prevented safe diving practices, a drop camera was used from the dinghy. The unit consisted of an underwater camera mounted on a cable, which relayed images to a monitor on the surface. Where shallow water depth prevented boat access, divers undertook swim-line search lanes to inspect these areas. Due to the presence of known terrestrial archaeological sites close to Konekaru Beach, an area up to 150 m off the beach and up to 100 m to either side of the mangroves was also inspected by divers using the swim-line search method. These surveys were later supplemented by a more intensive visual survey after the area had completely dried out at low spring tides. Mangroves and Inter-tidal Zone. The coastal fringe and intertidal zone was dominated by mangroves centrally located around the mouth of the Vaihui River, making survey particularly difficult. Archaeological surveys inside the mangroves were undertaken only within accessible areas of higher ground where the vegetation was sufficiently open for foot access. The higher ground was identified by two means. Firstly, the aerial photographs indicated the possibility of a small sand dune running through the mangroves. Secondly, potential for higher ground was inferred by reference to subtle changes in the vegetation. In both instances a survey team investigated the requisite areas to confirm both the presence/absence of higher ground and, where higher ground was identified, it was surveyed for archaeological material. All such sites discovered within the mangroves were recorded. 4.5.6 Background: Environmental Context 4.5.6.1 Climate During the summer months of December to April, the island of New Guinea witnesses the lowest barometric pressures of any island in the tropical Pacific. During this period, Papua New Guinea is subject to an extensive low pressure belt which extends from northern Australia to Indonesia and encourages monsoonal air flows from Southeast Asia. After passing the equator, these winds change direction and develop off the coast as the northwesterly monsoon. By July, the entire area is under the influence of the high pressure systems of the southeasterly trade winds. The periods of change in May and November are characterised by periods of doldrums and also considerable atmospheric instability. The southeast trade winds are predominant in the months of June to September, with the northwestern monsoons being experienced mainly from December to April (Darby, 1945:84-87). Oceanic currents during these periods roughly align with the predominant trade winds, with southeasterly currents experienced in August to September and northwesterly streams from February to March (Lewis, 1994:142-43). B. David, B. Duncan, M. Leavesley 4-479 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project These prevailing winds and oceanic currents are known to have significantly affected the scheduling of voyages for both local sailing ships and European masted vessels, and directly influenced the viability and staging of voyages carrying trade ceramics from the Boera area towards the Gulf Province in the past (the hiri trade, see below) (see David et al., 2001:70). 4.5.6.1.1 Water (Rainfall and Ground Water) The broader Port Moresby region including the Study Area represents the driest region of Papua New Guinea. ‘Permanent water supply’, writes Susan Bulmer (1978:13), ‘is one environmental factor that is likely to have been crucial to the location and success of [Koita and Motu] settlements’. Douglas and Sawanga (2007:22) write that ‘Very little is known about groundwater’ in the area, and ‘Boera residents obtain some of their water from local shallow groundwater wells. During the War the Gun Battery at Boera also developed a water supply for soldiers manning the gun battery by shallow bores, located by dowsing’. A resistivity survey undertaken in the 1980s at the junction of Papa-Leala and Porebada roads indicated ‘low volumes of poor quality water at shallow depths’ (Douglas and Sawanga, 2007:22). The Port Moresby region, including the Study Area, averages 995 mm per year (figures taken from Konedobu weather station, after Douglas and Sawanga, 2007:11). Rainfall is highly seasonal, with 76% falling between December and April (Figure 4.5.12). However, annual rainfall typically ranges from below 1000 mm to 1500 mm, and the region is subject to periodic drought, especially when wet season rains fail to arrive. 250 200 150 mm 100 50 A ug u S ep st te m be r O ct ob N er ov em D ber ec em be r Ju ly Ju ne ay M pr il A Ja nu ar F y eb ru ar y M ar ch 0 Figure 4.5.12 Mean annual rainfall for the Port Moresby area as taken from the Konedobu weather station (1881-1941), after Douglas and Sawanga (2007:11). 4.5.6.1.2 Temperatures The Port Moresby area sees little variation in mean monthly temperatures through the year, as common also of other tropical regions (Figure 4.5.13). B. David, B. Duncan, M. Leavesley 4-480 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 28 27.5 27 26.5 o C 26 25.5 25 ug u S ep st te m be r O ct ob N er ov em be D ec r em be r A Ju ly Ju ne ay M pr il A Ja nu ar y F eb ru ar y M ar ch 24.5 Figure 4.5.13 Mean monthly temperatures for the Port Moresby area as taken from Jackson’s Airport, after Douglas and Sawanga (2007:14). 4.5.6.2 Geology Much of the LNG Facilities site consists of flat coastal plains and tidal flats (see section 4.5.6.3 and section 4.5.6.6 for discussions of the shallow marine environment). Further inland, some 6 km from the coast, is ‘a range of hilly ridges running North-West to South-East, with wide valley floors between ridges … Those hills that exist near the coast [in particular near Boera] are lower, more rounded and more eroded’ (Douglas and Sawanga, 2007:17-18). Sediments are of marine origin, with sandstone, siltstone, limestone and chert being found across the region. The low-lying valley floors and coastal plains consist of recent alluvial deposits, with mud and clay being common in near-coastal and valley settings (usually on raised marine sediments), and skeletal soils in the hills (Douglas and Sawanga, 2007:18). Sea levels have prograded through the course of the mid to late Holocene, with the coastline presently being further to sea than in the past. Consequently, we can expect ancient coastlines dating to ancient settlements to have been further inland than presently, although precisely where these coastlines would have been at different times in the past remains a matter of speculation as systematic research has not been undertaken on this question for the Study Area (but see Betitis and Sullivan, 1990). Attempting to trace ancient shorelines by tracking through contour levels appropriate to past sea levels is not adequate for this region as significant amounts of alluviation can be expected as thousands of years of forest clearance and agricultural practices inland can be expected to have caused deep siltation near the coast through time. Proper appreciation of past cultural practices across the region will in future need to systematically address this question of the positioning of ancient coastlines. 4.5.6.3 Coastal Geomorphology and Pilotage The southern PNG coast from Kerema to Samarai is characterised by gradual subsidence with outlying chains of the main mountain axis not far from the coast. This has led to a terrestrial landscape of undulating hills, peninsulas and swampy plains. Rivers have partly silted estuaries and are generally shorter than those further to the west in the Gulf Province. This limited deposition of river silts has allowed more continuous development of barrier and fringing reefs in this area. The lower basin embayments of the larger rivers have resulted in mangrove fringes and swamps. Many sandy beaches exist along this coastline, particularly around Redscar Head, although lesser development of beaches has occurred in the areas down to Port Moresby as a result of the fronting reefs. Fringing and barrier reefs extend southeastward from Caution Bay all the way to south of Port Moresby and around Hood Point to Samarai. Caution Bay is an open cove, of limited use to shipping (due to its exposure to B. David, B. Duncan, M. Leavesley 4-481 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project prevailing northwesterly winds) and contrast with the safe harbour available at Port Moresby. The Port Moresby hills run out into the sea, and fringing reefs are located offshore (Darby, 1945:52-53). Navigation through the western approach to Port Moresby via Liljeblad passage has long been discouraged by local sailing directions, as it was reported to be ‘partly obstructed by foul ground and only suitable for vessels with local knowledge’ (Darby, 1945:234-5). 4.5.6.4 Vegetation Much of the Study Area consists of grassland (mainly Themeda australis and kunai grass), with Eucalyptus savannah occurring further inland and extensive but narrow stretches of mangrove along the coast (Douglas and Sawanga, 2007:25, 34-35). Bulmer (1978:19) writes that ‘the savannah and grassland vegetation [of the Port Moresby area, including the Study Area] are thought to be predominantly anthropogenic (Eden 1974)’, although Allen (1977:421) notes that there is ‘no clear consensus of opinion as to whether these grasslands are the result of low rainfall or burning by man’ and ‘the dry climate and highly seasonal rainfall of the Port Moresby area contribute to the ease with which the forest is cleared and the maintenance of the grassland through seasonal fires’ (Bulmer, 1978:16). This dry climate and periodic droughts gave rise to annual food shortages in the past: ‘oral traditions and historical records show that the recent [Koita and Motu] inhabitants of the Port Moresby area regularly had shortages of garden produce and depended upon imported food for much of the year’ (Bulmer, 1978:18) (see section 4.5.6 for further details). 4.5.6.5 Fauna Douglas and Sawanga (2007:28) note that ‘No data exists on the invertebrate fauna of the area’, and that many of the vertebrate populations have been affected and in some cases decimated by high human populations associated with PNG’s capital city Port Moresby. Despite an absence of systematic faunal surveys, eight species of snakes, about 70 species of birds, three species of native rodents, seven marsupials, pigs and bats are found in the general area (Douglas and Sawanga, 2007:26-31). While wallabies were once abundant, Douglas and Sawanga (2007:30) note that they are now ‘unlikely to be found due to hunting pressure from nearby villages’. Dogs appear to be a late Holocene import into this part of PNG: ‘In lowlands sites dog is not present in pre-ceramic deposits. … Dog was present in the earliest layer at Taurama (Site AJA), probably dating to about 2,000 years ago’ (Bulmer, 1978:30). 4.5.6.6 Shallow Marine Environments Douglas and Sawanga (2007:31) write that ‘The great barrier reef runs parallel with the Konebada coastline, petering out towards Galley Reach in the north west of the project area. The reef provides an important habitat for fish and marine invertebrates’. We have presented in section 4.5.5.2 details of reefs identified by local villagers in and near the Study Area. Here marine invertebrates (e.g. shellfish, prawns), fish and turtles are abundant and serve an important dietary role to Lea Lea, Papa, Boera and Porebada villagers (see also Douglas and Sawanga 2007:32-33). 4.5.6.7 Effects of Visibility and Substrate on Site Recovery Dense grass cover is well documented to inhibit ground visibility and site recovery (cf. Richards, 2008). The dense grass cover of much of the eastern part of the Study Area in particular can thus be expected to affect archaeological site recovery rates; this issue is addressed in section 4.5.7 below and has relevance to the cultural heritage sites surveys presented in this report. The environmental characteristics of the seabed in the Study Area have implications for the location and preservation of cultural heritage sites. In particular, the type of seabed substrate will determine whether or not sites will be preserved or able to be seen. For instance, a site will be well-preserved if it can sink into the seabed and be covered by sand or mud, as under such conditions it will have been largely protected from the effects of marine organisms such as teredo worms and adverse weather. Such conditions also provide good anaerobic (i.e. oxygen-free) environments for the preservation of a B. David, B. Duncan, M. Leavesley 4-482 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project range of materials including organics (such as wooden objects), metals, ceramics and glass. In such cases and if not disturbed, archaeological materials will remain relatively stable over very long periods of time spanning thousands of years (Dean et al., 1996). Organic materials, which usually deteriorate rapidly on terrestrial sites, survive extremely well underwater if buried, as evident in the discovery of 350 year old timbers from Dutch shipwrecks (Green, 1977) and prehistoric artefacts in the English Channel and European river beds (e.g. Fischer, 1995; see Van de Noort, 2008 for a general discussion of wetland archaeology). Seabeds with soft substrates offer the best potential for archaeological preservation, and therefore any archaeological site discovered in such conditions may well be of high archaeological significance (especially if organic objects are preserved). In environments with hard substrates (such as rock, reef or coral), archaeological materials tend to be quickly destroyed or broken apart, as such environmental conditions lack the protective covering offered through burial by soft sediments. Where hard substrate areas are also high-energy sites (e.g. surf zones), the water is also usually super-oxygenated, further accelerating corrosion or organic breakdown. Ships which wreck on hard substrates usually deteriorate and scatter at a much faster rate than those which are wholly or partially buried in softer, low-energy seabeds (Riley, 1997). Similarly rapid site break-down can be expected to apply to other types of archaeological sites, such as submerged villages, under such conditions. Variations in site conditions as outlined above (i.e. hard vs. soft substrate; low- vs. high-energy zones) also have implications for site visibility on the seabed (i.e. the potential to visually observe archaeological materials). Remote sensing techniques such as side-scan sonar and magnetometers may supplement visual inspection and aid in the relocation of submerged archaeological sites, particularly where water turbidity, full or partial covering in silts or sand, or depth restricts the ability to visually inspect sites. However, there are many factors and environmental circumstances which also limit or impede the success of such remote sensing techniques (see below). Much of the shallow marine environment in the Study Area consists of shallow mud and sand seabed overlying a hard compacted coral and rock substrate. This area extends up to 1 km offshore to where the edge of the shallow fringing reef, interspersed with coral and rock outcrops, drops to a deeper offshore channel. The seabed in this shallow marine area is unlikely to contain largely intact buried archaeological sites (such as ship and plane wrecks), as there appears to be insufficient depth in the substrate for cultural materials to be protected by soft sediments. However, it could hold the remains of Indigenous sites such as stilt villages and their associated structures or portable objects. The seabed in the deeper areas of the Study Area consists mainly of a sandy and/or muddy bottom of indeterminate substrate depth with greater potential for buried archaeological remains of all kinds. B. David, B. Duncan, M. Leavesley 4-483 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.7 Background: Cultural/Historical Context In order to meaningfully characterize and assess the significance of cultural heritage sites in the Study Area, those sites need to be understood in relation to the region’s cultural history. Furthermore, a review of this history may itself reveal the location of cultural heritage sites within or near the Study Area. There are five major avenues of investigation by which such an understanding may be achieved: archaeology (scientific research in historical places); archives (historical records including texts, photographs, drawings, audio-visuals); oral traditions (mainly as told by descendent communities); linguistics (language analyses); and social anthropology (scientific study of present-day and recent cultural practices). It is the aim of this section to investigate these sources for cultural heritage details that give information on the location, nature and significance of people’s past relations with place and the formation of cultural heritage sites in and adjacent to the Study Area. 4.5.7.1 Introduction to Ethnography (by Bruno David and James Weiner) The people of the Study Area today and in the past identify themselves either as Koita (or Koitabu as the Motu call them) or Motu, each group tracing their own, distinctive social and cultural history through ancestral male and/or female lineages. While the Koita identify their traditional lands as incorporating the Study Area’s coastal lowlands inland into the hills (see Figure 4.5.14), the Motu are restricted to a narrow coastal strip from Gabagaba in the southeast to Manumanu in the northwest, a distance of some 105 km (Allen 1977a:419; Chatterton 1968:92). Both Motu and Koita people have accounts of the places of origin of their ancestors and stories of the routes these ancestors took to arrive at their present-day locations. Included in the accounts of these journeys is the identification of previous village sites. These old village sites have the status of valued cultural places if not ‘sacred sites’ in the conventional sense of that term. Some of these sites appear to have been inhabited in recent times, such as the ancestral Koita village site at Mt Darebo, where pottery fragments are numerous; others such as Daeroto and Aemakara were inhabited in the more recent past. Davage, the coastal site directly north of present-day Boera village, is a previous village site, as well as an area of early Motuan settlement in the project area (see section 4.5.7.1). Because of the Motu-Koita custom of burial of bodies in the village precinct, all these sites almost certainly contain human remains. Today, each village has a more or less clearly defined cemetery within the village itself. Many of these more recent graves possess elaborate gravestones and memorials. B. David, B. Duncan, M. Leavesley 4-484 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.14 Map of Port Moresby region showing location of Koiraian language family (including Koita, marked as ‘Ka’ on the map; and Koiari, marked as ‘K’) (from Dutton, 1969:27). In the context of anticipating a major development project like the LNG152 Facility, Motu and Koita people are in some instances now asserting proprietary relations to places on the basis of oral and documentary history that indicates their ancestors lived in previous villages at those places. 4.5.7.2 Koita and Motu The Motu subdivide themselves, and are usually subdivided by researchers, into Western Motu and Eastern Motu, the former containing the ethnographic villages of Manumanu, Lea Lea, Porebada, Hanuabada and Pari (Figure 4.5.15). However (Allen 1977:424; see for example Chatterton [1968:92] for local oral traditions about the distinctive ancestry and origins of Boera Motuans, who call themselves Apau), [three other] Motu-speaking villages, Boera, Tatana and Vabukori, while geographically within the Western Motu, are considered in this local classification to be different from either group … Several cultural differences do exist apart from minor dialectal ones between the Eastern and Western Motu: the former are not known historically to have directly sponsored the hiri trading expeditions …, and at contact Eastern Motu villages were built completely in the sea, while Western Motu villages were constructed on the intertidal zone or on beaches. Bulmer (1978:41) thus writes that on ecological grounds: Oram (1975:31f) divided the Port Moresby villages into a western group, i.e. the Motu-speaking villages from Manumanu to Pari …, an eastern group, i.e. the four Motu-speaking villages east of Bootless Inlet, and the Koita and Koiari settlements. This coincides with the subdivision of the Motu into eastern and western groups mainly on military grounds (Groves 1972), except that three villages in the western ecological group, Tatana, Vabukori, and Boera, were traditionally independent of the other western villages. The western Motu fished, gardened, hunted and traded with the Koita and Koiari inland and the Doura and Gabadi to the north and east. Gardening was important to them, and their main crops were yams, bananas, and coconuts. B. David, B. Duncan, M. Leavesley 4-485 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project We follow Allen (1977) in that as the villages of Boera, Tatana and Vabukori follow the Western Motu in village locations and participation on hiri expeditions, they are combined with the Western Motu in this report. Interestingly, only the villages of Tatana and Vabukori – two of the three atypical Western Motu villages – have been ethnographically documented to manufacture the ageva shell disc valuables used for trade, further indication of the somewhat distinctive cultural character of these Motu villages (Allen 1977:430). Linguistic studies identify the Koita and Koairi (speaking non-Austronesian languages) as being closely related historically and as having a common linguistic ancestry that is different to that of the Austronesian-speaking Motu (Dutton 1969) (see Figure 4.5.14). ‘The Koita occupied territory inland from the western Motu, while the Koiari lived inland from the eastern Motu’ (Bulmer 1978:43). Swadling (1977:42) states that ‘Koita tradition says that they once lived on the Sogeri Plateau before they migrated to Rouna and then to the coastal lowlands. Oram has recorded one genealogy of 19 generations which goes back to the time when the Koita moved away from the Koiari [with which they were previously affiliated] (Oram n.d.)’ (see Gadiki, 1972 for oral traditions about the migration of Koita from the inland Naoro village to Nebira in the hills of the coastal plains; Dutton 1969 for linguistic evidence of linguistic divergence and movements toward the coast; see also Hicks 1973). Based on genealogical reckoning, the emigration of the Koita from Rouna-Sogeri is believed to have taken place perhaps 200 years ago, while ‘The pottery found on traditional Koita sites also indicates that the Koita movement from the Sogeri area was recent’ (Swadling 1977:42). Swadling (1977:42) continues: From Koiari, Mountain Koiari and Koita stories it would seem that Koiari advancements into Koita land was one of the main reasons for the Koita movement to the coastal lowlands (Dutton 1969:33, 43-5, 59). This Koita movement was greatly encouraged by fear of Koiari sorcery and their poisoning of Koita wells. Figure 4.5.15 Map of Port Moresby region showing location of ethnographic villages around 1870 (from Bulmer, 1978:figure 3.1) B. David, B. Duncan, M. Leavesley 4-486 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project In line with such oral accounts of the coastward movement of Koita villages, Allen (1977:430-31) suggests that ‘at the time of [European] contact the inland settlements were fewer in number than the archaeological evidence suggests for the preceding centuries’. For the most part, those inland settlements appear to have been hamlets with 5-10 houses rather than larger villages. The houses in these hamlets, ‘while … on piles, were much lower to the ground’ than the Motu coastal houses, and ‘each village or hamlet appears to have had one or more houses constructed for defence in the branches of tall trees, sometimes 15 metres or more above the ground’ (Allen 1977:431-32) (see Plate 4.5.2). Allen (1977:432) further notes that ‘While much of the stated Koita domain would be infertile hills, certain edaphically determined localities such as the fringes of the Waigani swamp and the stream and river banks would have provided areas of relatively good gardening land’. Plate 4.5.2 Left: ‘Sadāra Makāra, Koiari village near Bootless Inlet’. Right: ‘Tree house, Koiari village’ (from Lindt, 1887:plates XII and XIV). Western Motu origin stories tell of the coming of people from the east (Oram 1975), although some stories also tell of Motuans coming with pottery from the west and settling in villages to the immediate west of Port Moresby, including Lea Lea (e.g. Gadiki 1972:30-31). ‘The far western Motu villages, Rearea [Lea Lea] and beyond, were … better endowed than the more central villages’ of the Port Moresby area, ‘with richer fishing grounds, access to estuarine environment at Galley Reach and to garden land on the river. … explanations for the movement from the east into the central area … include warfare, population pressure and epidemic disease. The movement from the central area to the west can be seen as expanding to gain access to better resources on the part of groups forced into an unbenevolent situation’ (Bulmer 1978:41). Oram (1975:33-35) suggested that a population of 4000-5000 people lived in an area of around 500 km2 in the broader Port Moresby area in the 1870s; Allen’s (1975) review of the historical literature indicates a total of 3000-3500 people in the western villages of Manumanu, Lea Lea, Porebada, Hanuabada (consisting of an amalgamation of three Motu and two Koita villages), Pari, Boera, Vabukori and Tatana (see Table 4.5.1 for a village-by-village breakdown). Oram (1975:35, cited in Bulmer, 1978:47) suggests a total of around 1500 for the Koita. Bulmer (1978:48), following Turner (1877:474), stresses that these numbers are almost certainly lower than earlier population sizes due to the depopulation effects of European-introduced epidemic diseases such as measles, smallpox, ague, dystentry and whooping cough. B. David, B. Duncan, M. Leavesley 4-487 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.1 Population within the western Motu villages in the 1880s (excludes exclusively Koita villages) (after Allen, 1977:table 1, based on data in LMS Annual Report, 1890; Rosensteil, 1953:145; as see also Bulmer, 1978:47) Village Population Manumanu 300 Lea Lea 209 Porebada 349 Hanuabada 1310 Pari 306 Boera 315 Vabukori 184 Tatana 205 Total 3178 Oram (1970:7) notes that ‘In 1874, land held by the Motu was confined mainly to a narrow strip between the hills and the sea while the Koita held rights to large areas of land which extended inland as far as the Laloki River’. The Koita ‘acted as middlemen in trade between the Motu and the people of the mountains, exchanging Motu pottery, coconuts, fish and shell, for example, with meat, feathers, stone implements, garden produce and other commodities’ (Bulmer 1978:43). The Western Motu annually relied on their Koita neighbours for food during times of drought and famine in particular, for while the Motu obtained marine resources such as fish, crabs and shellfish, and to some degree produced yams, bananas and other garden produce, invariably the wet season brought resource shortages necessitating importation of foods through trade (see Allen 1977, 1991 for detailed discussions). ‘Motu have never been able entirely to live off the land within the narrow strip of relatively infertile coast that most of them inhabit’, writes Groves (1960:5). Long-term survival was thus achieved through short-distance exchanges with Koita and other neighbouring groups for garden and meat (in particular wallaby) produce, as well as through annual long-distance maritime hiri expeditions that brought back large quantities of sago from the Gulf Province (Plate 4.5.3). Ceramics, along with shell valuables and marine products such as fish, were the principle trade goods enabling the Motu to obtain such food resources (e.g. Seligmann 1910). As Groves (1960:7) points out, ‘Lacking dependable primary resources adequate to their needs, the Motu have founded their trading system on manufactured goods, especially pottery’. Through chains of connection, these ceramics and other trade products would find their way throughout the coastal and hinterland regions of the Central and Gulf Provinces and beyond (Plate 4.5.4). As Lindt (1887:124-25) wrote in 1887 about traded shell valuables: An article of very great value to the native is the ornamental toea or arm-shell. A few small ones are made on this part of the coast, but the best come from the east, as far away as the D’Entrecasteaux Group. They trade them for pottery, &, to the Dauni natives, whilst the Dauni natives sell them again to Mailuikolu for sago, dogs, &c., and these to the Aroma natives for pigs, dogs, and canoes. The Aroma natives trade them to the Hood Bay, Kerepunu, Kalo, Hula, Papaka, and Kamari natives for birds’ plumes of various kinds, and these again to the Motu natives for sago, and the Motuan to the Eelemaites for sago in bulk, weighing 2 or 3 cwt’. It is such objects found across the southern PNG landscape that has enabled archaeologists to shed light on this part of the country’s cultural history. B. David, B. Duncan, M. Leavesley 4-488 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.3 Left: sago being produced in Gulf Province. Right: sago bundles ready for local trade, Gulf Province. Both photos taken in 2007 (courtesy of Bernard Sanderre) Plate 4.5.4 Left: ‘Women’s canoe laden with pottery’ (from Chalmers, 1895:91). Right: ‘Women of Tupuselei, going for water’ (from Lindt, 1887:plate XXIV; this photo appears to have been the original for Chalmers’ drawing) Cultural heritage site types in the Port Moresby region are varied and include clay sources (Plate 4.6.6), freshwater wells (see Bulmer 1978:44 for Boera) (Plate 4.5.6), chert sources, ochre sources, and food resource areas (marine resource zones, hunting grounds, and Motu and Koita gardens; Plate 4.5.7) as well as villages (Plate 4.5.8) and burial places. Known clay sources include site ABG near Boera, and ‘two places 100 m apart on the beach between Rearea [Lea Lea] and Papa (Groves 1960:11) (Site AFC). The last sources were used by the people of Manumanu, Boera and Porebada’ (see section 4.5.7.9, section 4.5.7). B. David, B. Duncan, M. Leavesley 4-489 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.5 Woman digging for pottery clay near Porebada (photograph courtesy of PNG National Museum and Art Gallery: photograph 19695) Plate 4.5.6 ‘Well inland from old ridge site and present river mouth settlement of Kido’ (photograph courtesy of PNG National Museum and Art Gallery) B. David, B. Duncan, M. Leavesley 4-490 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.7 Gardens shortly inland of Boera, January 1974 (photograph by Gabrielle Johnston and Pamela Swadling, courtesy of PNG National Museum and Art Gallery) Plate 4.5.8 ‘”At low water,” native houses at Koilapu’ (from Lindt, 1887:plate XX) 4.5.7.3 Koita and Motu Settlements th Following Seligmann (1910:41), Bulmer (1978:45-46) notes that ‘In the mid- and late 19 century there were at least twenty-five separate settlements’ in the area between Bootless Inlet and Galley Reach, from the coast inland to the Laloki River (Figure 4.5.15). Ten of these were Motuan (some of which also contained Koita sections), while another 15 were Koita villages. Bulmer (1978:45-46) continues: All of the Motu villages and five of the Koita villages were located on a beach, either in the intertidal zone and/or above high water [e.g. Plates 4.5.8 and 4.5.9]. Three Motu villages were apparently completely on dry land; Manumanu and Rearea (Lea Lea), which were on beaches lacking the protection of a barrier reef (Groves 1972:803), and Pari (Chalmers and Gill 1885:258). … The Koita settlements that were not on the beach, were on hills behind the beach or some distance inland (Seligman[n] 1910:45; Dutton 1969:26-31). These settlements were in the process of progressive movement toward the coast at the beginning of the historic period. B. David, B. Duncan, M. Leavesley 4-491 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.9 ‘Women making pottery’ in a Port Moresby village (from Lindt, 1887:plate V). Allen (1977:427) notes that Western Motu houses ‘were located along the beach and intertidal zones … Houses were erected in lines according to iduhu groupings [social sections], all on piles at an approximate height of 3 metres above the ground for the shoreward houses’, and that ‘no Motu village was built higher than the beachline’ (Allen 1977:428). Bulmer (1978:46) states that other, special activity sites were used contemporaneously with these settlements, including ‘fishing and travellers camps on offshore islands, hunting camps inland, and garden camps in valleys in the coastal hills’. She further writes that ‘The Motu villages were spaced out along the coast at approximately 15 km intervals, with an extra distance between the eastern villages and the western villages, which formed military alliances in opposition to each other (Groves 1972:803). Separate Koita settlements were located either in between Motu villages or on elevated land inland from a Motu village’ (Figure 4.5.15). Bulmer (1978:49-55) undertook a detailed study of Western Motu and Koita settlements, taking into consideration their location, size and internal structures. We present her results in some detail as these will help inform the expected nature of village distributions, sizes and structures within and near the Study Area. Beach and offshore islet settlements and specialized sites include: • Villages with up to 60 or more houses aligned along the beach and facing inland. Burials were located ‘in front of (inland from) the dwellings, in graves in the sand in front of each house’. Lea Lea was of this form. • Villages located on the intertidal zone with houses over the water. Porebada contained 50 houses, while Boera contained some houses over water and some on shore. • ‘Fishing and shellfishing camps on offshore islands, particularly Daugo and Haidan islands’ (Bulmer, 1978:50). • Garden houses. Inland settlements and sites consist of: B. David, B. Duncan, M. Leavesley 4-492 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project • Small numbers of houses – typically one to three – in garden areas or hills. These were invariably Koita (or to the east Koiari) houses. • ‘Hamlets of four or five houses. These occurred either as individual hamlets … or as a nucleated settlement’ (Bulmer 1978:50). • Villages on ridgetops. ‘Both Koita [in the west] and Koiari [in the east] had settlements with two parallel rows of houses facing each other along a ridge … The houses were elevated on the slopes at the back and at ground level on the ridge at the front. A tree house was located at either end of the settlement for defense’ (Bulmer 1978:50; see also Seligmann 1910:44). • ‘Another variant of inland villages were the double line Koita villages … Two lines of houses faced each other across an open space, and burials were in the street in front of the houses. Seligman[n] generalized that average Koita villages contained 20-22 houses (1910:44), referring to two beach villages’ (Bulmer 1978:50). • Hunting camps. ‘Koita hunters spent weeks at a time inland in the vicinity of the Laloki’ (Bulmer, 1978:51). • Garden houses ‘in valleys in the coastal hills, where people normally resident on the coast lived during garden work’ (Bulmer 1978:51). For generations, Motuans have largely built coastal villages on stilts over water (Western Motuans typically over shallow water, in particular the inter-tidal zone, e.g. Plate 4.5.10; Eastern Motuans over deeper water, e.g. Figure 4.5.11). In more recent years, Motuans have started building residences on land, usually with associated small garden plots adjacent to dwellings. This contrasts to Koita houses, traditionally built on communally owned land belonging to the clan (Brammel n.d.). Plate 4.5.10 Stilt houses at Hanuabada, date unknown (photograph courtesy of Biama Kanasa Collection, University of PNG) B. David, B. Duncan, M. Leavesley 4-493 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.11 ‘Tupuselei, marine village (from the shore)’ (from Lindt, 1887:plate XXII) In the immediate vicinity of the Study Area, three ethnographically documented villages have been noted by Seligmann (1910), Kauga (2008) and Rakatani (2008): Dirora (a.k.a. Namura), Aemakara and Konekaru. Many other important ancestral villages known from oral traditions or historical records exist in the region, as outlined in this report, but the above two appear to be those closest to the Study Area proper. Seligmann (1910:41) thus writes that the Koita village of Namura ‘stood between Boera and Lealea in the bush, a short distance from the coast’ (Figure 4.5.16). Writing on 4th March 2008 on behalf of the Namura clan, Rakatani (2008:2), notes that the original name of Namura village was Dirora, and that it was ‘the biggest of all the villages ever known in [Koita] history’. Dirora, a village site important to the Namura clan, lies outside but close to the Study Area proper, but Portion 152 ‘is a parcel of land which includes several portions the Namura Mata clan ancestors have accessed to land usuage for hunting and gardening activities’ (Rakatani 2008:9). The ancestral village site of Dirora remains to this day of utmost significance to the Koita, and we have been told by Boera and Lea Lea villagers in particular that this location must not be damaged by the proposed LNG152 Facility developments. We return to this site in section 4.5.13. B. David, B. Duncan, M. Leavesley 4-494 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.16 ‘Map of coastal portion of Central District’, showing locations of village sites Namura (Dirora), Aemakara and Konekaru (from Seligmann, 1910:figure 3) Aemakara is an ancient Koita village mentioned by Seligmann (1910:figure 3, 41) and Rakatani (2008). Rakatani (2008:5) notes that ‘Perhaps around the mid 1750s [as indicated by genealogical reasoning], the Namura tribe broke camp at Boiodubu Darovaina and relocated at Mageto land. Rest of the family members moved together with the Isu tribe and established Aemakara village inland of Boera within the Laba land area (152)’. Gau Ario of Papa village identified Aemakara as an important ancestral village during interviews at Papa on 25th January 2008. The ancient village of Konekaru was occupied at the time of initial European contact (e.g. Seligmann 1910:41). It is an important ancestral place of oral tradition that formed a gateway to the coast for the Koita (Rakatani 2008:11), and that enabled Koita and Motu fishers to co-ordinate fishing and sailing activities (Kauga 2008:2). Konekaru village and its associated beach are located immediately to the north of the Priority Area, and must not be damaged during developments. We shall return to this site in section 4.5.7 and section 4.5.13 below. The present villages of Lea Lea and Porebada may have been settled less than 200 years ago by genealogical reckoning, although archaeological research may reveal earlier cultural deposits not evident in oral traditions. According to these oral traditions (Swadling 1977:40-41; see also Hicks 1973; Oram 1968:87-89; Tau 1976), About 6 to 4 generations ago [in 1977], people began to leave Badihagwa and establish the village group now known as Hanuabada. Tanobada was built below the mission at Metoreia, Elevala around the island of the same name, and Poreporena on the mainland further to the east (Oram n.d.). Later people left Hanuabada and founded four new Motu villages and a section settled at Boera. The new villages were B. David, B. Duncan, M. Leavesley 4-495 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project LeaLea, Pari, ManuManu and Porebada. The last village, Porebada, was founded not long before the first Europeans landed in the early 1870s (Oram n.d.). 4.5.7.4 Social Structure (by James Weiner) The social structure of these villages appears to have been predominantly based on agnatic descent groups (iduhu) with both ascribed and achieved leadership. That is, descent-group headmen coexisted with ‘big-men’ who achieved renown through their economic and political exploits. Residentially, iduhu constituted separate parts of any village area. Groves observed that ‘there is no traditional government of any formal kind at the village level’ (1963:17) and that mobilization for collective action always occurred at the instigation of particular local patrilineal corporate groups or iduhu (Groves 1963:17). 4.5.7.5 The Men’s House or Dubu (by James Weiner) Rosenstiel1 wrote that the ‘chief’ supervised the activities of the iduhu in regards to activities such as gardening, marital problems and land tenure (Rosenstiel 1953:14). He was also the owner or tauna of the right front post of the dubu, or ceremonial platform. The dubu itself consisted of four such carved posts ‘each of which was hereditary within the iduhu which had built it’ (Rosenstiel 1953:14). Though she is not clear about this, Rosenstiel seems to infer that the dubu represented at least four separate iduhu, to one of which the ‘chief’ belonged and whose iduhu was represented by the right front post. Belshaw (1957:15) later wrote that ‘each iduhu has the right to a pattern of carving, which may be used on a post erected by any of its members’. He goes on to say that ‘it usually happens that a dubu is erected by representatives of several iduhu and often of several villages, each striving to outdo the rest in the provision of food for the feasts that accompany the stages of erection, and the final dancing and ceremonial’ (Belshaw 1957:16). ‘In no sense may the dubu be regarded as the exclusive property of one iduhu, though it may be said to be erected in honour of the iduhu of the originator; and it is a long-lasting symbol of an event in which the co-operation of many groups has been obtained’ (Belshaw 1957:16). In unpublished papers, Nigel Oram (Oram Papers NLA, Box 2, Folder 14) wrote that ‘If the Iduhu got too big it would break away and build a new Dubu. The name has nothing to do with family names but explains why some Iduhu have the same name’. None of the villages in the Project area have dubu (there is one in Hanuabada) and they do not comprise part of iduhu functioning today. Nevertheless, it is possible that dubu may be found archaeologically within the Study Area, and as indicated by the ethnographic evidence would be recognisable through the presence of postholes for a front platform on the first house along a row of houses. Such archaeological evidence would be significant in that it would enable a historicising of the iduhu socioorganisational system itself (see section 4.5.7.6 below). 4.5.7.6 The Iduhu (by James Weiner) The iduhu is the unit of social recruitment in Motu-Koita society. Belshaw (1957:12-13) saw it in spatial terms: ‘It consists of one or more lines of houses built on piles over the sea at an angle to the coastline by people who give themselves an iduhu name’. He goes on: ‘The iduhu … is primarily a residence unity based upon one or more separate lineages of patrilineal emphasis, and hence may be differentiated from a clan, which, in a technical sense, must consist of people claiming common descent’ (Belshaw 1957:13). James Weiner (2008) found this characterization to be accurate for the iduhu which he came into contact with during fieldwork in 2008. Groves (1954:78) likewise wrote in 1954 that: 1 Rosensteil probably acquired much of her information on the dubu from pages 60-65 of Seligmann (1910). B. David, B. Duncan, M. Leavesley 4-496 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Traditionally, the members of each iduhu resided in a single line of houses running out from the beach into the sea; the iduhu was therefore a discrete residential unit within a wider residential unit. Membership of an iduhu was determined patrilineally … Observers such as Groves, Oram and Goddard maintain that the iduhu recruitment principle is flexible and can incorporate non-agnates (non-patrilineally-related individuals). Rosenstiel similarly draws attention to the flexibility of iduhu structure when she describes that ‘… families were loosely organized into exogamous patrilineal groups called iduhu’ (Rosenstiel 1953:14). Goddard (2001:314) writes that ‘I believe iduhu have been more resilient than previous researchers have thought, that the importance of landholding has been underestimated in previous attempts to understand what iduhu are’. Groves also wrote ‘the iduhu is a corporate group of great importance’ (Groves 1963:17). He stated that during his fieldwork period, ‘conflict within the village is generally viewed as conflict between component iduhu’ (Groves 1963:18). Oram wrote in 1970 that ‘Iduhu were corporate groups with assets which included insignia, a name, fishing or hunting nets, and frequently but not invariably land’ (Oram 1970:7). He went on to say that ‘In most villages, land was divided into sectors and each iduhu regarded a particular sector as its own iduhu land’ (Oram, 1970:7). Iduhu land did not constitute a discrete territory either: Oram (1970:7) wrote that: Not all land … was iduhu land. Iduhu as corporate bodies did not necessarily hold rights to all land inside their land boundaries and lineages within an iduhu may have established exclusive rights to particular areas. Oram wrote that among the Koita, a man could confer land to his daughter, which would then belong to her and her descendants (Oram 1970:7). Oram wrote in his notes that women could assert rights to use their descent group land as well as men: ‘In the past rights to land were essentially rights to use the land when there was a need to do so. It appears that rights were enjoyed equally between male and female members of a descent group. The control of the land, however, rested with the senior member of the male line’ (16th July 1963: Oram Papers NLA Box 1, Folder 6). But he also wrote elsewhere (on 29th December 1963) that Gavera Baru told him that ‘the land belonged to the father or brother and his descendants, not those of the women’ (Oram Papers, NLA Box 1, Folder 6). Groves (1963:21) wrote that despite the fact that non-agnates can under certain circumstances be incorporated into the iduhu, the normative principle governing iduhu composition is that of agnation, or patrilineal descent. Therefore, ‘rights in a man’s estate are normally shared by his sons. Ideally, then, primary membership in an iduhu should coincide with membership in an agnatic lineage’ (Groves 1963:21). Oram also noted that at least one informant stressed the corporate nature of descent group land ownership: ‘Phillip emphasized that in the past there was no land belonging to a person but it belonged to the iduhu’ (Oram Papers, NLA, Box 1, Folder 6). As a result of movement, migration and warfare, iduhu have fissioned (and sometimes fused) and iduhu of the same name can be found in different villages, often with an additional name to differentiate them from other local iduhu of the same original name (Groves 1963:16; Goddard, 2001:315). Goddard (2001:315), writing more recently, observed that ‘The Motu-Koitabu themselves now translate “iduhu” (referring to groups within villages) into English as “clan”, reflecting the influence of common, rather than academic, terminology and an idiom that iduhu are principally patrilineal descent groups’. Groves observed that if a local iduhu (what he called a ‘section’) divides into two, the two resulting groups become separate and independent in every meaningful political sense, although they will continue to recognize ties of kinship between them (Groves 1963:16). The following can all contribute to the definition of iduhu membership and identity: residential unit, possession of a ceremonial platform or verandah, and exogamy. In 1954, Groves (1954:80) mentioned B. David, B. Duncan, M. Leavesley 4-497 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project ‘Tenure of a house and of garden land, property in ceremonial and ritual observances, and the cooperative labour of other members of the iduhu, were all very important interests which the unity of the iduhu safeguarded’. Groves made this characterization of iduhu functions in 1963: a name, a heavy fishing net, ability and equipment with which to build trading canoes, residential locality, claims to certain ceremonial dances, ritual relationships with identified deities, responsibility for amassing bridewealth, responsibility for distributing food at ceremonial feasts, possession of garden land, a leader, and a corporate history (Groves 1963:18). Groves also wrote that ‘The iduhu is the unit which owns and controls the assets that are most important to the Motu, such as fishing nets, trading vessels, ancestral rites, and, in some cases, land’ (Groves 1963:17-18). Rosenstiel wrote that the iduhu ‘controlled its own land, and garden plots were divided among the men, who owned the land’ (Rosenstiel 1953:15). As Goddard points out, however, all these attributes, while they may contribute to iduhu identity, are not primary in the way land-ownership is, in his opinion (Goddard 2001:316). However, Groves denied that all iduhu were land-owning units, nor were all landowning units iduhu (Groves 1963:26), and noticed that cognatic descendants could claim control of land in the absence of patrilineal heirs (Groves 1963:26-27). A term used by Groves which is also relevant to the task of describing agnatic solidarity within an iduhu is siahu which he translates as ‘potency’ (Groves 1963:23). ‘Siahu’ can also mean ‘spirit’ (commonly in the context of church related translations). Goddard more recently has glossed siahu as ‘power, potency’ and also ‘heat’ (Goddard n.d.). ‘Heat’ in this context refers to the potency of ancestral power performed in specific places and the legitimacy this lends in land claims. Goddard (n.d.:6) describes ‘heat’ in the context of making claims to land in the following way: Genealogical connections to specific ancestors legitimated narratives of the past, which might include, for example, stories of the movement of ancestors from place to place establishing or abandoning villages or gardens, fighting battles, killing or being killed and buried. Through these narratives speakers would iterate their siahu, or that of their iduhu, to inhabit, or use, or pursue various activities at, the places to which they referred. In other words, their siahu derived from their ancestor’s presence and actions at a given place. Any archaeological evidence that would shed light into the history of iduhu – such as structural evidence from dubu, as noted above – would thus have considerable significance, for iduhu are ethnographically of notable importance in social and territorial structure. 4.5.7.7 Settlement Patterns in Historical Perspective Bulmer (1978:51-54) has modeled population migrations and shifting settlements in the broader Port Moresby area from oral histories largely collected by Nigel Oram in the 1960s and 1970s. She writes that: Motu villages were based on two movements of people into the area, one from an inland area to the east to Motupore and Taurama, and another from the west in the Gulf of Papua to Boera. These were possibly ‘migrations’ in the sense that they consisted of people deliberately moving to a new territory, but it could be that they only involved small groups of people. Both these movements are from areas where other languages than Motu are now spoken [Figure 4.5.17]. Oral histories … indicate that by perhaps 450 years ago there were three villages on the coast of the survey area [broader Port Moresby area], i.e. Boera, Vabukori, and Taurama (Oram 1975:6), founded by the above movements of people, and another [further to the east of Port Moresby] … Since the period of the earliest four villages there has been a rapid increase in the number of Motu settlements, according to the oral histories. The settlement of Taurama was abandoned as a result of th warfare, probably in the 18 century, and by about 200 years ago there were additional settlements at Pari, Manugava (an inland hilltop site), Badihagwa (near Hanuabada), as well as at Vabukori. In the following century four other western Motu villages were founded to the west of Badihagwa, and three B. David, B. Duncan, M. Leavesley 4-498 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project others were settled to the east from Tubusereia. It can be suspected that this increase in numbers of settlements reflected a rapid increase in population … Allen (1975:9, 13) argues that pottery manufacturing centres along the Papuan coast appear to coincide with the largest population centres and that there was a development of a Motu ‘central place’ at Hanuabada in response to the Motu role as specialist traders. It seems to me that this increase in the th th number of settlements in the 18 and 19 centuries reflects instead a decentralization of Motu communities in general … During the period of Motu settlement and expansion, the Koita were in the process of moving from the inland river plains to the coast. … The histories of the Koita communities on the coast recount the th progressive movement from site to site toward the coastal hills, where they were settled in the 19 century. There was also some Koita resident in all but one Motu village at the beginning of the historic period, but th none of the separate Koita settlements were on the coast until the end of the 19 century. The Koita still maintained land rights as far as the Laloki River, using the inland area for gardening and hunting and collecting. Figure 4.5.17 Migration histories according to oral traditions (after Oram, 1981:214). Settlement locations across the broader Port Moresby region, including the Study Area, were dynamic. Warfare and migrations (due to conflict, resource shortages and the like) led to resettlement, sometimes in entirely new areas, and the ‘number of settlements has changed in the proto-historic period from four to fourteen Motu settlements … [while there was] probably an increase in the number of Koita settlements’ (Bulmer 1978:55). The Koita initially had ‘small inland hamlet settlement’ (Bulmer 1978:55), while Motu settlements were coastal or over water, but through time each attracted each other into more or less symbiotic relations as the Koita moved towards the coast and in later times came to share coastal villages with the Western Motu. In the 19th century, villages tended to contain a few hundred people, typically 200 to 350 (see Table 4.5.1), although this may be an underestimate for earlier times due to the unknown effects of 19th century diseases. Settlements typically ‘ranged in size from a single house to over 60 grouped together. The smaller settlements were Koita, but there were also larger Koita villages, of a similar size to the Motu villages’; there was ‘considerable variation in the arrangement of nucleated settlements’; and both ‘Koita and Motu had limited activity sites, for fishing, hunting, gardening, which were occupied temporarily for special purposes’ (Bulmer 1978:54). It is salient to remember that we cannot expect oral traditions to have retained all knowledge of all times in the past, especially in such a region where settlement systems are known to have been historically so active. Rather, much of the oral history relates to the last 500 years (or less), and for B. David, B. Duncan, M. Leavesley 4-499 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project any given period of time some village and other cultural sites can be expected to have been forgotten, especially the further one goes back in time. It is this kind of forgotten site that archaeological research can reveal, despite their absence from oral testimonies (along with forgotten historical details of remembered sites). It is also significant to note that the largely settlement-subsistence concerns of most if not all archaeologists working along the southern PNG lowlands in the 1960s into the 1980s, when all of the region’s professional archaeological research was undertaken, took place at the expense of ritual, sacred and religious sites. The outcome is an impoverished and almost entirely ignored archaeological record that relates to such latter cultural concerns (a common feature of pre1980s archaeological research in many parts of the world). 4.5.7.8 Burial Practices Little is known of Koita and Motu burial practices. We cite Allen (1977:432), who summarises well the existing evidence: Direct evidence of Koita burial practices is poor, and this probably means that the Koita sharing Motu villages followed the Motu custom of interment in shallow graves within the village, with either a small grave house above, or a stake on which the weapons or utensils of the deceased might be placed. The grave would be filled with beach sand, and might be disturbed at a later date to recover various bones to be worn by relatives. Inland Koita practices may have been more like Koiari, where primary interment seems to have been less favoured than exposure in houses, on platforms, and bundle burials in trees. Secondary burial took place, and in one account a pit grave is described as circular, lined with stones and having a small wooden conical frame above on which was suspended the valuables of the deceased. Archaeological evidence would add cave and rock fissure interment to these practices. Such burial practices would express themselves archaeologically as interments within or near house spaces within hamlets and villages; inland as ethnographically or locally known burial platforms or burial trees; and mortuary caves and rockshelters. 4.5.7.9 The Hiri Trade The peoples of the Port Moresby area – in particular the Motu but also to a much lesser degree the Koita – were renowned makers of ceramic vessels. ‘All of the Motu villages made pots, with the exception of two, Vabukori and Tatana, that specialized in the manufacture of shell ornaments … Thus there were manufacturing specialties even among the villages participating in the same trade system’, writes Bulmer (1978:42, following Oram 1975); ‘The potters were described as recent immigrants to the area, one group coming from the east to Taurama, and another group coming from the west to Boera’. During ethnographic times the pottery-making villages included Porebada, Boera, Lea Lea, Manumanu, Tatana, Pari, Hanuabada, Elevara and Tanabada (Lampert 1968:77, after Barton 1910; Chalmers 1887:11; Haddon 1894:149). Clay sources have been documented ‘on the coast between Lea Lea and Papa (Groves 1960:11), Tubusereia, Boera, and Pari … that were used in the protohistoric period by Motu potters. These all seem to occur in the coastal areas’, although ‘inland clays were probably used’ (Bulmer 1978:15). Pottery was manufactured by women both for domestic use and for local, regional and distant (hiri) trade. The regional trade involved women carrying pots by canoe or on foot to kin or trade partners in nearby inland Gabadi, Doura and Koita villages (in particular villages along the Aroa River), in exchange for garden and meat produce, in particular yams and bananas. In time the Gabadi, Doura and Koita villagers themselves would exchange some of these pots further afield, resulting in a widespread spatial patterning of ceramic pots amenable to archaeological investigation (Groves 1960:8). The hiri (see Dutton 1980 for its linguistic history) is an ethnographically reported trade system involving Austronesian-speaking (principally Western Motu) ceramic pot manufacturers and traders sailing annually to villages in the Gulf of Papua (Plate 4.5.12). The hiri trade journeys are well documented in the late 19th and early 20th century literature (e.g. Barton 1910; Chalmers 1895; Chester 1878; see Oram 1982 for a review). Trade voyagers set-off in fleets of (typically around 20) B. David, B. Duncan, M. Leavesley 4-500 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project multi-hulled sailing ships (lagatoi) from the Port Moresby area of Bootless Bay and Caution Bay (including Boera-Papa-Lea Lea; e.g. Anon. 1929) when the southeast trade winds blew, typically in October or November, and returned with the monsoons around January. These trading expeditions brought ceramic pots and shell artefacts to the western Gulf Province villages, in return for sago and canoe hulls that would be strapped to the ships for the return voyage. So large were these expeditions that Seymor Fort (1886:15) wrote in his government report in 1886 that annually ‘20,000 pots were taken, for which they would bring back in exchange about 150 tons of sago’; other estimates indicate around 30,000 pots and 600 tons of sago per annum (e.g. see Allen 1977b; Allen and Rye 1982 for reviews). Motu traders regularly travelled to the Gulf Province coastal villages as far west as Vaimuru along the Purari River delta, and there are suggestions in local oral traditions that the Motu trade expeditions sometimes went further west (e.g. Chester 1878:9) (Plates 4.5.13 and 4.5.14). These villages then served as redistribution centres for inland villages and villages further to the west (e.g. those of the Kikori River and nearby river systems) (e.g. Chester 1878:9; Oram 1982). The finding of a rock painting of a large, lagatoi-like crab-claw canoe on Dauan in northern Torres Strait (Brady 2006; McNiven et al. 2004:244) suggests that at least on rare occasions hiri traders may have ventured even further west to northern Torres Strait, whether by accident or design. As Groves (1960:8) concludes from the ethnography, the Motu hiri trading network was ‘more extensive than any other yet reported from Papua and New Guinea’, and in this holds a special place in PNG’s cultural history. B. David, B. Duncan, M. Leavesley 4-501 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.12 Lagatoi on Port Moresby Harbour, date unknown (photograph courtesy of Biama Kanasa Collection, University of PNG) Plate 4.5.13 ‘Fleet of lakatois starting for the west’ (from Chalmers, 1895:75) B. David, B. Duncan, M. Leavesley 4-502 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.14 ‘East End lakatoi at Purari Delta’, moored off Kaimare village around Christmas 1915 (photograph by Ernest Sterne Usher, courtesy of South Australian Museum, AA835 C93; see also Pike and Craig 1999:234, 248. From David, 2008b:figure 1) Professional archaeological research since the late 1960s indicates that the ethnographically recognisable hiri system and its associated ceramic traditions probably began around 500 years ago (see David 2008). Older ceramic traditions across the Gulf and Central Provinces also suggest that the historical hiri descended from a further 1500 years or more of formalised long-distance maritime trade relations across the region (e.g. Allen 1976, 1977; Bulmer 1978, 1982; Rhoads 1982; for a review and significantly expanded radiocarbon chronology, see David 2008) (see section 4.5.8.5 for details). At the other end of the chronological spectrum, hiri expeditions were severely disrupted during World War II when Motu villages were evacuated and also as a result of increasing involvement in the wage economy since the mid-1900s (Ryan 1970; see also Johnston 1974, cited in May and Tuckson 2000:59). Formal hiri trade expeditions continued sporadically into the 1960s, although they largely ceased in the late 1950s following the sinking of a lagatoi off the coast of Boera village in 1957 (a then-predominant Motu hiri pottery manufacturing centre), when several lives were lost (the location of this sunken lagatoi off Boera has not been mapped, but its presence should be kept in mind by the developers). However, long-held trading partnerships between villages have in many cases been maintained, despite the demise of formal hiri expeditions (Vincent Eka, Kerema resident and descendent of historical hiri trade partner, personal communication 2007). The Origin of the Hiri Motuan hiri trading vessels and expeditions are well described by early observers (e.g. Barton 1910). Lennox (1903) describes an 1883 expedition in this way: These Motuans are the traders of Eastern New Guinea. The staple manufacture of the district is pottery, and the earthenware vessels made by the Motu tribe are used for cooking and other purposes throughout the land. The generic name for articles of this ware is uro; but uro is really the cooking vessel, while water vessels, dishes for serving food, large and small cups, small pots, large and small basins, pots with rims, and large vessels for holding sago are varied forms of domestic utensils manufactured by the Motuans, and each has its particular name. The distribution of uros is secured by barter. Food-stuffs are brought into Port Moresby and exchanged for uros, or the trading Motuan voyages along the coast and barters his uros for other commodities. Once a year the Motuans make a trip of two hundred miles to the westwards, faring forth with boat-loads of pottery and – in more recent years – of knives, beads, looking-glasses, red cloth, and tobacco; purchase in exchange large quantities of sago; and sell that again to the coast natives nearer home, receiving payment this time in arm-shells and other articles that represent the native currency. B. David, B. Duncan, M. Leavesley 4-503 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project This great westward trip is made by a fleet of lakatois, vessels made up by the combination of several large canoes, and capable of carrying a considerable crew and a large cargo. Here is Tamate’s description of these strange craft: ‘Four large canoes are lashed together. Their bulwarks are made from the leaves of the Nipa palm sewn together, well fastened with long, strong mangrove poles, and caulked with dried banana leaves. A stage is made all round, so that the sailors can work her without getting inside of the bulwarks. Masts of mangrove, with the roots, are stepped on to the centre, and large sails, made of mats all sewn together and shaped like crab toes, are fixed for working, with ropes made from the bark of the large yellow hibiscus. The anchor is a large stone made fast with long canes, sometimes one hundred fathoms in length. Fore and aft are small covered-in houses, strong enough to withstand a very heavy sea, where the captain, mates, and boatswains sleep and smoke. There are strong divisions of wicker work in each canoe, into which pottery is put, each division having an owner. The pottery is well packed with dried banana leaves, and only when thrown ashore in a gale do they have much breakage’. …On this occasion Tamate secured a passage on board the Kevaubada, one of these lakatois, and, after a voyage of five days, arrived in far-distant Elema, making the port of Vailala. The Kevaubada was a twomaster, and he took up his sleeping quarters on two planks covered with a mat and set on the top of a large crate of pottery between the masts. [See Chalmers 1895:74-92 for a first-hand account of this hiri expedition]. Motuan oral tradition has it that the hiri trading voyages originated at Boera village (e.g. Barton 1910; Lewis 1994:134-35). In 1910, F. R. Barton published a widely known origin story in Seligmann’s The Melanesians of British New Guinea. This oral tradition has been handed down for generations and continues to be retold today by Western Motuans of the Study Area. We recount the origin story here in some detail because of its great significance to Motu history, and because it relates directly to a number of cultural heritage sites located near the proposed Portion 152 Facility site (see section 4.5.7). These cultural heritage sites, by virtue of their direct association to this hiri origin story, are of utmost cultural, historical and educational significance – and central to the maintenance of cultural traditions by local peoples – and must therefore not be damaged as a result of the proposed LNG152 developments. We will return to the mapping of these sites, and their management recommendations, in section 4.5.7 and section 4.5.13 respectively. ‘A very long time ago’, writes Barton (1910:97-100), ‘there lived at the Motu village of Boera a man named Edai Siabo’. He continues: One day he sailed with some other men in a canoe to the islands of Bava and Idiha (small coral islands on the barrier reef off Boera) to catch turtle. They were unsuccessful, and at night the other men went to sleep on the island, whilst Edai Siabo, who was varo biaguna (‘master’ of the turtle net) slept alone in the canoe. During the night a being named Edai, of the kind called dirava, arose from the water, seizing hold of him and carrying him under water to the cave among the rocks which was his abode. The dirava drew Edai Siabo head-foremost into the cave so that he lay prone with his feet projecting from the entrance, and he then informed him that he had brought him there to tell him about lakatoi (composite trading canoes). ‘Do not be afraid,’ he said; ‘as soon as I have told you all about lakatoi, you can go back to your canoe.’ The dirava went on to explain how these vessels should be made, and how, if he and his fellows went to the west in a lakatoi, they would be able to obtain plenty of sago to die them over the season of scarcity. At daylight next morning the men who had slept ashore swam out to the canoe, and when they saw that Edai Siabo was gone they wept. While they were talking, and weeping, and wondering what had become of him, one of them looked over the side and saw their comrade’s feet and called to the others to come and see. So they all dived into the sea and caught hold of his feet, and tried to haul him out of the cave, but the dirava held the shoulders of Edai Siabo, and the men could not move him, and they had to rise to the surface again to take breath. Again and again they dived down but were unable to pull him out for the dirava still held fast to Edai Siabo because he had not finished telling him about lakatoi. At last, when all had been told he allowed the men to haul Edai Siabo out of the cave to the surface of the sea, and they placed him in the canoe. He was apparently dead and the men wept sorely over him, but after a while he B. David, B. Duncan, M. Leavesley 4-504 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project opened his eyes and revived. His companions asked him what he had been doing, and he told them that he had seen and heard many strange things. When the men asked him what these things were, he told them that the dirava Edai had taken him into his rock-cave, and instructed him as to the manner of making a lakatoi, and about the hiri (the trading voyage on which the lakatoi must sail). The men inquired the meaning of these words, and Edai Siabo promised that he would repeat all that the dirava Edai had said to him when they had returned to Boera. So they made sail for that place. There Edai Siabo built a model of a lakatoi according to all that the spirit had told him, and when he put it upon the sea it sailed along quickly, and all the assembled people exclaimed: ‘Inai! (behold!) who taught you to make such a thing?’ and he told them that the dirava Edai had taught him thus to make a big vessel, and to sail in it to the west for sago. Then he took the little lakatoi to his house, and the men of the village went there to examine it and ask questions. Edai Siabo explained to them how to lash the canoes together, and how to step the mast, and how to make the sail, and so forth. So the people went away and built a lakatoi, and they called it Oalabada. A Koita – a brother-in-law of Edai Siabo – tried to dissuade him from going to the west, telling him that in his garden there were plenty of bananas, and in his house good store of yams, so that he would not want, but Edai Siabo remained stubborn. When the lakatoi was finished it was loaded with earthenware pots, and as soon as all the pots had been stowed aboard the people wanted to dance on the lakatoi, and they called for their drums; but Edai Siabo forbade them to beat drums on the vessel. He told them that instead of drums they must use sede (a percussion instrument made of bamboo), and he explained to them how these should be made. So the men went into the jungle, and cut bamboos and made sede, and when they beat them they were delighted with the sound given forth. After that they went aboard again, and poled the lakatoi through the shallow water, intoning meanwhile the following words: ‘Dokaimu Oalabada dokaimu, Ido-Ido, Ido-ido-ido-ido,’ and all the while they kept beating the sede. Presently they asked what song they should sing, and Edai Siabo then told them the words and tune of the lakatoi ehona (song) as the dirava Edai had taught him, and the words of it were these: ‘Oalabada Oviria nanaia Ario Visiu O Veri Auko Bogebada Eraroia Nanaia Irope Umanai Ela Dauko’ (and many other verses). When the song was ended those who were not going on the hiri went ashore, and the others hoisted the sail and left. They sailed for many days into the west until they came to a large village on the banks of a river, and there they stopped. The people received them with great joy inasmuch as they never before had pots in which to boil their sago. The travelers remained there until all the pots had been bartered for sago and then the lakatoi being loaded they set sail for home. Now Edai Siabo was married to a woman named Oiooio, and when he sailed away to the west, he told her that after fifty days were past, her daughter-in-law was to climb every day to the summit of the hill called Taubarau, to look out for the lakatoi returning. Day after day she returned to Oiooio saying she could see nothing. The wives of the men who had gone, took other husbands, but Oiooio remained faithful, in the sure belief that her husband would return, till one morning her daughter-in-law said she had seen something near Varivari islets, but she could not be sure that it was not a piece of floating driftwood. Oiooio told her to hurry back and look again. As it came nearer and grew larger she saw it was indeed the lakatoi and ran down to tell the good news. Oiooio swept the house, washed herself, put oil upon her body and in her richest ornaments paddled off to the lakatoi when it rounded the point to the village. There she told those aboard that their wives had been faithless, and that she and her daughter-in-law had alone been obedient to the commands imposed on them by Edai Siabo before leaving. She took some sago from the lakatoi and returned to her house, and after Edai Siabo had washed in the sea, he and those with him went ashore. The men were greatly grieved to find that Oiooio had spoken the truth about their wives, for many of them were big with child by other men. Then Edai Siabo told all the people that the words of the B. David, B. Duncan, M. Leavesley 4-505 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project dirava were all true, and he admonished the faithless women and the men who had taken them as their wives. The women were very ashamed of themselves, and some of them were taken back by their husbands. Since that time the lakatoi have gone every year to the west, and there has consequently been food in plenty during the season of scarcity. 4.5.7.10 The Ceramic Industry Murray Groves (1960:3) writes that in the 1950s ‘Motu pottery traditionally found its way, and still finds its way, into almost every village along the shores of the Papuan Gulf and in the immediate hinterland’ (e.g. Plate 4.5.15). The ubiquity of this cultural product gives it great archaeological potential, allowing archaeologists to investigate cultural change, including past inter-regional relations and interactions across close and distant communities. Plate 4.5.15 Ceramic pot (uro) in the Gulf Province village of Epemeavo in August 2007, obtained in the past through hiri trade The late 19th and early 20th century ethnographic records from Motu pottery manufacturing villages identify a number of formal pottery shapes and decorative designs within a single general ceramic style. Pottery was made in most Motu-speaking villages (including Delena village near Yule Island to the west of the Study Area, where Motuans are said to have lived in the past). Seneca (1976:4) describes how a Boera woman called Boio Siabo introduced Western Motu knowledge of pottery manufacture to the Koita after she was ‘carried off by a Koitabu tribesman called Bokina Bokina after a tribal war raid. She spread the knowledge of pot making to her husband’s village women’. Following Groves (1960), Haddon (1894:156) and Stone (1880:141), Bulmer (1978:55-56) thus notes that ‘Pottery is also made in Koita-speaking communities … but it was generally thought that the Koita learned the skill from the Motu’. Numerically predominant among ceramic vessels were uro cooking pots, hodu water jars (typically larger and deeper than the uro) and nau dishes (Arifin 1990:31). As Arifin (1990:31-39) notes, however, other named forms were also present (such as kibokibo, e.g. Bulmer 1971), with Chalmers B. David, B. Duncan, M. Leavesley 4-506 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project (1887:122) documenting 10 named vessel types, Finsch (1914:270) eight, and Barton (1910:114) seven; more recent, mid-20th century commentators have documented up to 12 Motu pottery types (Table 4.5.2). Not all of these pottery types are said to have been traded by the Motu. Furthermore, a number of pot shapes were further sub-divided into size classes by the Motu to create a broader range of distinctive and recognised vessel types (Arifin 1990:35). Plates 4.5.15 and 4.5.16 and Figure 4.5.18 show examples of vessel types recognized by the Motu during ethnographic times. Plate 4.5.16 Motu ceramic pot types from Manumanu, February 1974. Left: tohe. Middle: hodu. Right: nau. ‘Man in photo called Karai; his mother made the pot (tohe) in the time of his grandmother. Urimu his mother died about 1970’ (photograph courtesy of PNG National Museum and Art Gallery) Figure 4.5.18 Motu ceramic pot types (from Bulmer, 1978:figure 3.2) B. David, B. Duncan, M. Leavesley 4-507 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Pottery was manufactured by the Motu and, to a lesser extent, the Koita for a number of reasons: domestic use, short-distance (mainly inland) trade with the Gabadi, Doura, Koita and Koiari (Bulmer 1978:56), and long-distance (maritime) hiri trade with Gulf Province communities. The pottery was made with paddle and anvil technique (rather than coil technique as practiced in some other parts of Melanesia) (Plate 4.5.17), the paddles commonly being ridged, although ‘This ridging is normally erased by the potter in the final paddling with a smooth paddle’ (Bulmer 1978:57) (Plate 4.5.18). Ceramic manufacturers made both plain (undecorated) and decorated wares, the latter representing makers’ marks enabling the male traders to keep track of whose (female kin) products they were exchanging (see Groves 1960 for details of such siaisiai services). However uro, in ethnographic times the principle trade item, were usually undecorated. More generally, pottery made for domestic use was undecorated (Bulmer 1978:61). Bulmer (1978:57, 59) notes that ‘pottery decoration has been rapidly forgotten’ by recent Motu and Koita generations, and ‘the historic period has seen the reduction of the “kinds” manufactured from ten to four, only one of which remains numerically common’. Plate 4.5.17 Pottery being manufactured at Porebada (probably in the mid-1970s) (photograph courtesy of PNG National Museum and Art Gallery: photograph 19696) B. David, B. Duncan, M. Leavesley 4-508 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.18 Ethnographic paddles and anvil for pottery making, collected from several households in Manumanu in 1974 (photograph courtesy of PNG National Museum and Art Gallery). It is also widely recognized that ceramic traditions have changed significantly through time. Bulmer (1978:59-60) thus notes that: The distribution of pottery making is said to have changed during the proto-historic period, with separate introductions of pottery making from both the west and east into an area for which no earlier tradition is described. The fact that Motu style pottery was found to be made in two non-Motu-speaking settlements to the east and west may be taken as a possible indication of a process of expansion of the industry in the proto-historic period. Indeed, Haddon (1900:275) said that pottery making was introduced into the Yule Island area by the Motu. Another change in the pottery industry in the historic period has been the reduction in the number of pottery-making villages, and in the quantities of pots. Previous archaeological research both within the Central Province and Gulf Province has revealed the existence of a range of past ceramic conventions that were not practiced during ethnographic times: ceramic conventions have changed significantly through time. This historical dynamism highlights the significance of archaeological ceramics as testimony to past ceramic-making traditions (in a way that oral traditions alone cannot due to loss of such details from social memory) across the Motu homeland, including the Study Area. The best way to adequately reveal this ceramic history is to adequately sample each site in order to understand the antiquity and geographical spread of specific ceramic conventions. The clay and sand temper used in the manufacture of pottery will also differ from village to village, and through time. ‘In spite of the basic common technology [the widespread use of paddle and anvil technique]’, writes Bulmer (1978:57), ‘the differences in clays, tempers, and the individual trademarks ought to provide a basis for identifying the villages of origin of Motu pottery’ (e.g. Plate 4.5.19). Each archaeological site is a unique historical ‘document’ possessing its own, singular evidence of past cultural activity (and in many cases village sites will themselves have been more or less specialized pottery manufacturing centres, the history of which tracks changing spatial configurations of ceramic manufacture, specialization and trade). Lampert (1968:77) thus concludes: The dominance of Motu pottery and its widespread distribution through trade makes its ethnographic and archaeological study a vital one, not only for the history of the Port Moresby district but for that of a large part of Papua. … Largely on the basis of pottery analysis we can reasonably expect B. David, B. Duncan, M. Leavesley 4-509 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project archaeological sites in the Port Moresby district to provide a sequence of material culture reflecting both the movements and identity of people … the sites will no doubt reveal long forgotten and unrecorded facets of the everyday lives of people. Table 4.5.2 Function Large cooking pot Traditional categories of Motu pottery (after Bulmer, 1971:63, 1978:58; Groves, 1960:14). Stone (1876) ura – 15-18”Ø Lindt (1887) uro – large vessel Finsch (1903) uro – everted rimmed spherical pot Small cooking pot keikei – small pot Sago storage pot tohe ? kaeva – pot with rim kaiwa – pot with horizontal ‘collar’ Seligmann (1910) uro – 10-12”Ø Groves (1960) uro – 10-16”Ø everted rimmed spherical pot keikei – small pot shaped like uro Water vessel hordo hodu – water vessel hodu – spherical pot with narrow neck and vertical rim Serving dish nao nau nau – oblong dish with lugs at either end ? ohoru – large cup oburo – deep slightly incurved bowl ? ituru – small cup itulu – cup with goblet-like stem and base ? kebo - basin ? kibokibo – small basin tohe – same shape as uro, but several times larger tohe – same shape as uro, but several times larger hodu – 12-18”Ø hodu – 12-18”Ø nau – 12-20”Ø circular bowl nau – 12-20”Ø circular open dish itulu – small basin with legs, for dye kebo kibo – basin It is clear that any disturbance or destruction of archaeological sites in and near the Study Area needs to systematically sample all threatened sites prior to any disturbance to ensure adequate characterization of ceramic conventions by which this part of PNG’s history can be understood and recorded, for once destroyed such sites can never again be accessed, and such historical details would be gone forever. It is clear that any disturbance or destruction of archaeological sites in and near the Study Area needs to systematically sample all threatened sites prior to any disturbance to ensure adequate characterization of ceramic conventions by which this part of PNG’s history can be understood and recorded, for once destroyed such sites can never again be accessed, and such historical details would be gone forever. 4.5.8 Archaeology: Indigenous Cultural Heritage Austronesian-speaking villages (such as the Motu) are found scattered along the eastern part of the southern PNG coastline, concentrated in the Central Province, and many of these manufactured pottery in the recent past. The implication is that ancestral Austronesian maritime colonizers – the ancestors of the Motu – began to settle in the region some 2000 years ago or perhaps earlier. Although earlier ceramic sites have not yet been found in the southern PNG lowlands, McNiven et al. B. David, B. Duncan, M. Leavesley 4-510 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project (2006) have argued that around 2600-2400 years ago Austronesian speakers came to Torres Strait likely from the east; they anticipate that ceramic sites of similar antiquity will one day be found along the ancient southern PNG coastline (located along the inland plains shortly to the north of the present coastline), for this is the most likely source of influence for the manufacture of the Torres Strait ceramics. Plate 4.5.19 Examples of ceramic sherds collected from different archaeological sites by the PNG National Museum and Art Gallery, by site code. Top left: AAJ. Top right: ACQ. Bottom left: AMG. Bottom right: AMH. The contrast in decorative conventions between sites is clear, and shows how different sites have the potential to reveal unique information on specific pottery conventions. Scales differ between sites (scales not included in the original photographs) (from photographs courtesy of the PNG National Museum and Art Gallery) The Port Moresby area has received professional archaeological attention since 1967. Here can be found a wide variety of ceramic decorative styles, and it is the history of these ceramic traditions that has formed the focus of archaeological research in this region (in part because of a focus on the history of the hiri trade and of earlier long-distance maritime trade networks which involved the manufacture of large quantities of trade ceramics over some 2000 years or more). In this section the history and key findings of archaeological research in the broader Port Moresby area are presented, as they inform an understanding of Lea Lea-Boera’s cultural heritage sites and their significance. Many of the sites discussed below possess their own language names (obtained from oral traditions or the vicinity from which they are found) (e.g., Nebira), archaeological code given by the discovering archaeologist as part of their own surveying referencing system (e.g. Nebira 2), B. David, B. Duncan, M. Leavesley 4-511 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project and/or a unique three or four letter reference (e.g. ACJ), being the official PNG National Museum and Art Gallery site register code (by convention, site lettering is organized by PNG Province; all registered cultural heritage sites from the Central Province and the National Capital District begin with the letter A). For example, the cultural heritage site known from oral traditions as Nebira has been sub-divided by archaeologists into a series of distinctive, archaeologically separate exposures each of which has been given a separate researcher reference number (e.g. Nebira 2, Nebira 4 etc.), and each of which has been given an official PNG National Museum and Art Gallery site number (Nebira 2 = ACJ; Nebira 4 = ACL). We emphasise that despite 40 years of research and numerous systematic and opportunistic cultural heritage site surveys across different parts of the broader Port Moresby region, only a small proportion of cultural heritage sites has so far been recorded. No archaeological surveys had previously been undertaken in the Study Area itself. For ease of report structure, the results of previous archaeological research are presented by researcher name, in general chronological order. 4.5.8.1 Graeme Pretty In 1967, Graeme Pretty undertook reconnaissance archaeological surveys in the vicinity of Boera village, in search of Maurice Leask’s (1943a) previously reported ‘kitchen midden’. Pretty undertook preliminary surveys on and around Stanley Hill, recording three sites (which he termed Sites A, B and C), but without finding the sought-after sites. He notes that ‘both the Summit and slopes were thickly strewn with potsherds, shell and other Melanesian habitation residue’ (Pretty 1967:34). During these investigations, Pretty visited Boera village and the nearby beach, recording in the process the important cultural heritage site of Edai Siabo’s first lagatoi anchor (Pretty 1967:35) (which he identifies as the anchor of the sailing ship by which Edai Siabo founded Boera; see section 4.5.7.9 for details of the legendary story of Edai Siabo and his first lagatoi). The anchor was already partly covered with sand at the time of Pretty’s visit. 4.5.8.2 Susan Bulmer Susan Bulmer’s 1978 doctoral thesis Preshitoric culture change in the Port Moresby region is the only PhD dissertation, and the largest single study, ever undertaken on the archaeology of the Port Moresby area. It represents the culmination of research she began in 1967, and supercedes many of the conclusions about the region’s archaeological past she presented in earlier publications (e.g. Bulmer 1969, 1971). Bulmer was interested in understanding the distribution, and ecological and social inter-relationships, of sites across the landscape, and how spatial variation and temporal change in ceramic conventions could be used to explore the region’s cultural and social history. She argues that settlement-subsistence systems shifted through the course of Port Moresby’s preEuropean contact history, and these changes were accompanied by shifts in the location of potteryproducing centres and changes to ceramic styles. She suggests that during the Early Period of occupation, from around A.D. 0 to 1000, a relatively homogeneous pottery style was widespread along the Central Province coast from Mailu in the east to Yule Island in the west. Towards the end of the Early Period, a large settlement could be found at Ranvetutu. During the Middle Period, from around A.D. 1000 to 1500, the earlier pottery style rapidly changed, making way for ceramic conventions akin to those of Milne Bay some 370 km to the southeast. Towards the commencement of this period large pottery-producing communities were set up at Motupore and Boera, while previously established communities at Taurama, Nebira and Eriama continued to exist. During this time, pottery-using settlements became established on elevated hills in the coastal hinterland, probably for reasons of defense. The Middle Period was followed by the Proto-historic Period (about A.D. 1500-1875) immediately preceding, and continuing into, the early European contact period, when ‘Middle period pottery is replaced by a single style, which in the 18th and 19th centuries appears to sub-divide into the eastern and western variants’ (Bulmer 1978:xxi). The late Proto-historic Period saw a predominance of settlement on the coastal hills and along the coast, and ‘heavy dependence upon imported food based on the specialist manufacture of shell ornaments and pottery, was of relatively B. David, B. Duncan, M. Leavesley 4-512 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project recent origin’ (Bulmer 1978:xx). Bulmer (1978, 1982) argues that the people of the ancestral Nebira, Eriama and Taurama villages – spanning nearly 2000 years of occupation – were not specialized craft manufacturers (for a different view, see Allen and Rye 1982; Allen et al. 1977), and that while there is evidence in oral traditions and in the archaeological record for close contacts between coastal and inland communities, these sites show little evidence of specialized trade (a point disputed by Jim Allen in particular – e.g. Allen and Rye 1982). Bulmer suggests that early in the region’s history large settlements containing ceramics were established on the inland river plains. During the last 300 years (based on oral traditions), she argues that settlements shifted towards the coast. She asks if the earlier, hinterland villages were occupied by the Koita (the ‘people of the land’, who possess the oral traditions about those older sites), while the later coastal settlements were occupied by the Motu ‘people of the sea and trade’ (sometimes together with the Koita). Using oral traditions and historical records, she interprets the archaeological evidence around the notion that the Koita ‘had moved down from the mountains and across the plains to the coast, while the Motu arrived by sea to dwell with them’, both movements taking place only during the last 400 to 500 years, with the Koita ‘reaching their position in or near Motu villages in the 19th century’ (Bulmer 1978:39). Yet the Koita did not traditionally practice pottery-making, having learnt the craft from the Motu after the latter’s arrival along the coast (perhaps 2000 years ago, perhaps more recently with earlier ceramic manufacturers in the Port Moresby region before the Motu). If the hinterland villages indeed relate to early Koita occupation, what of the pottery found within those sites? The ‘great variety’ of pottery shapes and decorative designs found across the Port Moresby region are critical to unraveling such questions, for it is the pottery that holds the key to understanding the location of manufacturing centres, the arrival of new pottery manufacturers, cultural affiliations between communities and geographical dispersal routes through trade and acculturation. By determining the location of pottery manufacturing centres (villages) through time, the geographical spread of ceramic traits across the landscape – and the occurance of specific styles at any given archaeological site – will allow understanding of landscape use and social relationships between sites across the broader region. Bulmer’s work on the history and dynamics of ceramic production and settlement location in the broader Port Moresby area was based on the analysis of pottery sherds collected from 67 archaeological sites within an area covering 800 km2, and the excavation of two ancient village sites and a rockshelter, Nebira 2 (ACJ), Eriama 1 (ACV) and Taurama (AJA). Her investigations focused on the region from Bootless Inlet in the east to Galley Reach in the west, from the coast northward to the Laloki River. Within this area the Koita and Motu have long lived in a ‘complementary relationship in an overlapping territory’ (Bulmer 1978:2) involving trade and cohabitation in close social relations. The archaeological ceramics of the Port Moresby region contain a range of vessel shapes and decorative designs, many of which are not represented by ceramic conventions of ethnohistoric times (cf. Plate 4.5.19). The geographical distribution and antiquity of each of these conventions is of great interest to archaeologists, for it is through these that the spatial, temporal, cultural and social relationships of sites can be investigated. As Bulmer (1978:74) notes, ‘pottery decoration is particularly sensitive to culture change’ (much like car designs today mark the changing times). Archaeological pottery remains have been the major means by which the region’s cultural history has been investigated, including patterns of inter-regional interaction and the history of the hiri system itself (e.g. Allen 1977). Consequently, every site containing pottery remains will hold archaeological significance. Here we summarise the major pottery decorative styles identified by Bulmer (1978) for the Port Moresby region (incorporating Lea Lea-Boera). We note that while the chronological value and spatial integrity of these styles remain in contention by archaeologists (e.g. Allen 1977; Swadling 1980), Bulmer’s schema remains one of only two detailed published accounts by which archaeologists presently order Port Moresby ceramics. And here-in lies a major problem: Bulmer’s ceramic styles are ordered into an apparently chronological system but are not, in themselves, based on systematic B. David, B. Duncan, M. Leavesley 4-513 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project temporal data. We thus note that while her ceramic schema requires refinement and perhaps replacement once fine-grained chronological investigations have taken place (Bulmer did not undertake fine-grained temporal analyses of the sites she excavated), her framework remains the archaeologically predominant one and is therefore cautiously presented here. Bulmer’s study is largely based on 2977 ceramic sherds from 67 undated surface archaeological sites (Bulmer 1978:76-77). Her six decorative styles are summarized in Table 4.5.3. Table 4.5.4 represents her style key for the region’s ceramic bowls. She argues that four cultural phases can be identified for the broader Port Moresby region based on changes in ceramic conventions, as indicated by her surface ceramics, combined with results from three archaeological excavations she undertook at Nebira 2 (ACJ), Eriama 1 (ACV) and Taurama (AJA) together with those of others (principally Motupore, Nebira 4, Ava Garau) (Bulmer, 1978:340-41): 1. Early Period with Style I pottery: around A.D. 0-1000; 2. Middle Period with Styles II, III and IV pottery: around A.D. 1000-1500; 3. Proto-historic Period with Styles V and VI pottery: around A.D. 1500-1875; 4. Historic Period: after around A.D. 1875. B. David, B. Duncan, M. Leavesley 4-514 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.3 Summary of some characteristics of decorative styles of Port Moresby bowls (from Bulmer, 1978:table 5.5). Style I Common Techniques Red Slip Vessel Forms Probable Associated Pot Decoration Characteristic Rim or Lip Form Slipping Burnishing Incising Combing, grooving Simple restricted bowl Simple unrestricted bowl Composite restricted bowl Composite unrestricted bowl Thickened, round Thickened, square Round Round Heavy line incising, perforation Appliqué Grooving IIa Composite bowl Square, round IIb Simple unrestricted bowl IIc simple restricted bowl Thickened, square Thickened, round Slipping Burnishing Incising Painting II Eriama Incised/Applique (formerly Massim) III Eriama Incised/Punctate (formerly Massim) Fine line incising, punctation Simple restricted bowl Thin, round IV Taurama Shell/Comb (formerly Boera/Taurama) Shell and comb impressing, combing Composite bowl Square Shell and comb impressing, painting V Taurama Incised/Punctate (formerly Motu) Heavy line incising Simple bowl Thickened round or square Incising VI Waigani Incising, finger impressing, shell impressing Simple bowl Thickened round or square B. David, B. Duncan, M. Leavesley 4-515 ? ? ? Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.4 1 2 3 4 5 6 7 8 9 Pottery style key for Port Moresby bowls (from Bulmer, 1978:table 5.6). Slipped or burnished Style I Not slipped or burnished Go to 2 Shouldered vessel with incised, perforated, punctuate, grooved or appliqué decoration Not so Style IIa Go to 3 Unrestricted vessel with wide, decorated horizontal lip Style IIb Not so Go to 4 Thickened lip Style IIc Not so Go to 5 Thin lip and lacking thickened rim Style III Not so Go to 6 Square or round lip and shell or comb decoration Style IV Not so Go to 7 Thin lip with thickened rim, and incised or punctuate decoration, or undecorated Not so Style V Go to 8 Vessel with A5 and/or A6 field decorated using Design Units #76 and 39 (see Bulmer, 1978) Not so Style VI Go to 9 Residue: ‘foreign’ sherds B. David, B. Duncan, M. Leavesley 4-516 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project While Bulmer (1978:96) notes that the earliest, Style I ceramics are found in the Lea Lea-Boera area, such early ceramics are concentrated along the coast but remain relatively uncommon across the Port Moresby region. Red slipped and/or burnished sherds of this early phase have long been documented from Central Province and Gulf Province archaeological sites (e.g. Sullivan and Sassoon 1987). Because of this previous identification of red slipped and/or burnished ceramics from radiocarbondated sites elsewhere in the PNG lowlands (e.g. Vanderwal 1973), Style I is the least controversial phase of her chronological sequence. Styles II, III, IV and V sherds are also found in the Boera area (Bulmer 1978:96, figure 5.17). If Bulmer is correct in identifying her styles as chronological markers, the implication is that the Lea Lea-Boera area contains archaeological deposits covering much of the period of human occupation in the Port Moresby region. Given the large number of archaeological sites identified in the Study Area and the proven presence of early ceramics in the Lea Lea-Boera area, we can thus expect the Study Area to contain cultural deposits spanning a considerable period of time, and potentially incorporating materials from Style 1 ceramics onwards. 4.5.8.3 Jim Allen Jim Allen’s work in the Port Moresby region involved both field research and the theoretical modeling of culture change in this ceramic manufacturing and ethnographically renowned long-distance maritime trading centre. Allen (e.g. 1984:415-16) notes that the Motu, like other southern lowlands PNG Austronesian-speaking groups, did not settle rich agricultural landscapes but rather coastal regions fronted by resource-rich offshore reefs. These were (and continue to be) specialized maritime peoples who also gardened, hunted and gathered, but it is the sea that formed the focus of subsistence and settlement practices. Nevertheless the drought-prone Port Moresby region, and the paucity of agricultural products directly available to the maritime specialist Motu, meant that alternative means of obtaining food resources had to be developed to ensure long-term survival. The answer came in the form of craft specialization (ceramics and shell valuables used for bride price and the like) and long-distance maritime trade (the hiri). Both these ceramics and shell valuables have high archaeological visibility enabling the history of such trade and social relations to be investigated. Jim Allen undertook archaeological excavations at two ancestral village sites in the Port Moresby region, Nebira 4 (ACL) and Motupore (AAK). Both sites contain rich cultural deposits, including two human burials, 7483 flaked stone artefacts (amongst which are obsidian pieces imported from Fergusson Island, and drill points), 49,728 pottery sherds, numerous animal bones (mainly pig, wallaby, fish and shell), shell artefacts (including beads and fragments of arm bands) and varied pieces of ochre and ground-stone artefacts from Nebira 4; and 40 burials, numerous stone drill bits, hundreds of shell disc beads, large volumes of shell and vertebrate faunal remains (particularly marine and wallaby), structural evidence in the form of pits and post holes, and very large quantities of ceramic sherds from more extensive archaeological excavations at Motupore (e.g. Allen 1977:443, 444; the Motupore results remain largely unpublished). One of these Motupore burials (a secondary burial with a dog’s teeth necklace) dated to around 400 years ago is interpreted as Koita, due to its similarity to Koita and Koiari burials of ethnographic times. The implication is that by that time KoitaMotu relations were already close enough for a Koita burial to be included in a predominantly Motu village, as was the case during ethnographic times (Allen 1977:445). Nebira 4 is believed to date from around 2000 years ago to sometime before the colonial period. Together with the similarly-dated Eriama 1 and Loloata Island midden, it currently represents the oldest known site in the PNG southern lowlands. The similarity of dating of the earliest cultural levels at each of these sites, along with Oposisi in the western Central Province where 2000 year old ceramics were also found, led Allen (1972:121) to conclude that ‘we appear to be dealing with a widespread maritime migration into the central coast about 2,000 years ago. These people established themselves widely and maintained good communications for at least a thousand years [as indicated by similarities of ceramic conventions between communities]’. B. David, B. Duncan, M. Leavesley 4-517 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project The Nebira 4 faunal assemblage indicates a marine-oriented economy (as would be expected of early Austronesian speakers such as the ancestral Motu) during the earliest (and thus oldest) cultural layers, becoming gradually less so through time (Allen [1972:116]; this may be due to increasing dependence on inland gardens, as Allen [1972:122] suggests, or to subsequent sedimentation of the coastal plains). The ceramic sequence indicates an early red slip (and sometimes burnished) tradition followed by a sequence of ceramic conventions including continuity of red slipping (Allen 1972:99). Allen (1972:105-9) identifies nine decorative styles (Styles A-I), many (but not all) of which represent sequential changes in ceramic conventions. The Nebira 4 sequence can be summarized into a sequence of three phases (Allen, 1972:108, 109): Horizon 1. Levels 1-8. Globular pots with heavily rolled horizontal rims; bowl forms shallow and open, often with a thickened lip; decorative style A the most distinctive marker, with a large percentage of painted pottery. [Corresponds with Styles IA and IB at Oposisi]. Horizon 2. Levels 9-15. Globular forms a mixture of horizontal and angled rims with the latter more popular; deeper bowls with straight sides; styles D and E the most common decorative styles with some temporal value, together with styles F and G. [Corresponds with Style IIA at Oposisi]. Horizon 3. Levels 16-19. Globular forms with angled rims; bowl forms most commonly restricted, and found in association with decorative styles F and G. Styles H and I are the best indicators of this early horizon. [Corresponds with Styles IIB and IIC at Oposisi]. The dating of these phases remains unclear due to dating uncertainties and insufficient radiocarbon determinations to resolve such questions (Allen 1972:121). Nevertheless, Nebira 4 clearly demonstrates some 2000 years of ceramic evolution and the deeply stratified potential of archaeological sites in the Port Moresby region. Motupore in Bootless Inlet to the southeast of Nebira was established around A.D. 1200, and appears to have been abandoned around A.D. 1700 (Allen 1977:443). Motupore is of salutary interest in that it is only referred to once in the recorded oral traditions of the greater Port Moresby area, yet as determined archaeologically it was once a major site of ancestral Motu character (Allen 1977:442, 446). The lesson is that even major archaeological sites may not register in oral traditions, forgotten from social memory through time. Allen (1984:420) wrote that Motu (and to a lesser degree Koita) pottery ‘underwrote the emergent maritime trading systems’. Allen (1977) has suggested that socioeconomic interactions between the Koita and Motu, and with trading partners further to the west in the Gulf Province, have intensified through time. Such intensifications are observable archaeologically in a simplification (decreased decoration) and standardization of Motu ceramics with the mass production of trade goods, along with an increased population evident in a concomitant proliferation of occupational sites. Among the Western Motu, amicable relations with the Koita led to the establishment of seaside villages, but further to the east less amicable relations between the Eastern Motu and the Koiari led to the construction of Motu villages over the sea for purposes of defense (Allen 1977:451) (e.g. see Plate 4.5.11). He notes that pottery-producing Motu settlements were located in low-rainfall parts of PNG subject to periodic droughts, encouraging the development of specialized pottery manufacture for which food products (in particular sago) could be traded in surplus quantities (Allen, 1984). Nevertheless the manufacture of (principally hiri) trade ceramics did not simply meet the dietary needs of the Motu villages, but also enabled the fulfiment of high risk, status-enhancing longdistance maritime voyages and the acquisition of surplus products (sago) by which internal exchange relations could develop (with Koita and other nearby groups). The development of specialized ceramic-for-food trade relations with long-distance trade partners (in the Gulf Province) as well as with neighbouring groups (such as the Koita and Koiari, the latter bringing shell lime and highlands stone axes to the Motu) created social developmental momentum that gave rise to the complex Motu and Koita societies of ethnographic times. The Study Area’s archaeological ceramics are thus not simply archaeological objects worthy of museum display, but more importantly allow an understanding of how B. David, B. Duncan, M. Leavesley 4-518 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Motu and Koita history and culture developed through social interaction. This rich ceramic tradition now evident in the region’s archaeological record has a long history whose details still largely remain to be revealed. Following Bulmer (1971), Allen (1977:439-42) initially divided Port Moresby’s archaeological sequences into three broad periods, which he referred to as the Early Ceramic Horizon (A.D. 0-1000), followed by a ‘middle period’ onto a ‘final period’ (a re-formulation of his Nebira 4 three-phase sequence, see above). He suggested that during this initial ceramic phase, Austronesian speakers came from the east and settled in an interconnected network of villages along the southern PNG coast, maintaining between themselves good inter-community communications and thereby a commonality of ceramic conventions. However, ‘The demise of this Early Ceramic Horizon is sudden all along the coast’ (Allen 1977:448). The subsequent phase of the ‘middle period’ saw ‘the possible removal of the people from the valley floor site of Nebira 4 to the adjacent hilltop site of Nebira 2 and the occupation of the offshore island site of Daugo near Port Moresby’ (Allen 1977:439-40). Allen here suggests that around A.D. 1000 the (presumably Austronesian-speaking) people of the Early Ceramic Horizon came under pressure from inland (ancestral Koita) groups as the latter began to move towards the coast, necessitating the establishment of settlements in more defensive positions (hilltops and offshore islands). Following Bulmer (1971), around A.D. 1000-1400 two new ceramic traditions then appeared in the Port Moresby area: intrusive (i.e. foreign) ‘Massim’ wares from the Milne Bay area, most evident from archaeological sites in the Boera area; and ‘Boera/Taurama’ wares that appear to represent ancestral Motu ceramics. The pottery of the ‘final period’ corresponds to the ethnographically recorded Motu ceramics. Allen (1977:446) suggests that as Motupore was occupied continuously from around A.D. 1200 to 1700, and as Motupore’s most ancient ceramic decorative styles can be shown to evolve uninterrupted into decorative conventions that are akin to Motu ethnographic examples, its inhabitants were likely ancestral to present-day Motuans. ‘For this reason a certain adjustment needs to be made to Sue Bulmer’s proposed culture sequence’ (Allen 1977:446), which posits a sequence of interrupted ceramic styles representing external influences or replacements. Hence, as the ceramic conventions of Bulmer’s ‘Boera/Taurama’ Middle Phase are found at Motupore, where they can be shown to be ancestral to, and evolving into, historic Motu incised/impressed wares, Allen (1977:446) suggests that the later two stages of Bulmer’s sequence should be coalesced into one, reducing the entire Port Moresby sequence into two phases: an early phase spanning around A.D. 0-1000; and a later phase beginning ‘somewhere before A.D. 1200 and continuing to present’ (Allen, 1977:446). Allen concludes that the long-debated ‘hiatus between the two is therefore reduced, and it is into this hiatus the Massim industry described by Bulmer must be fitted. The status of the people represented by this pottery still requires elaboration … On the present evidence it may well be that there was no hiatus at all, and that the Massim component infiltrated during the brief period of disequilibrium following the disappearance of the earlier inhabitants and during the establishment of ancestral Motuan groups’ (Allen 1977:446; see also Swadling 1976). Motupore’s most recent phase, beginning around A.D. 1200, has a ceramic industry that can be followed uninterrupted into ethnohistoric Motu ceramics. This phase is interpreted by Allen (1977:446) as indicating that the Motu ‘impinged upon the existing central Papuan coastal population from outside the research area some 800 years ago’. That is, around A.D. 1200 a new wave of Austronesian speakers came from the east to the Port Moresby area with new ceramic decorative conventions, establishing a base at Motupore. These were the ancestors of the ethnographic Motu. Through time, as the Motu established and consolidated their villages along the coast, the Motu proliferated on the coast and the Koita both inland and on the coast as the two groups entered into symbiotic social and economic relations (Allen, 1977:449). In his later work, Allen (1984:423) argued that craft specialization was ‘vitally important’ to the Western Motu (and Koita) trade economy, and that they were ‘the only notable producers of pottery along some 400 km of the south Papuan coast’. Of note is the highly standardized ceramics that emerged during this recent, monopolizing phase, which Allen (1984:423) associated with increasing commercialization of production. Following Groves (1960), Allen B. David, B. Duncan, M. Leavesley 4-519 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project (1984) noted that the hightened levels of trade generated by establishing trade partnerships led to increased (and surplus) food returns into the Motu villages, which in turn fed increasing trade relations with neighbouring groups who brought hinterland food products (garden produce, wallaby meat) for imported surplus sago and ceramics, positively feeding back to higher populations that enabled the system to grow. By the later stages of the recent phase, this demographic growth had led to further increasing demands on food resources that led the ceramic-manufacturing women to work ‘at breakneck speed’ to produce the very large quantities of pots necessary for exchange expectations, in particular in the form of the long-distance hiri expeditions; ‘insufficient care in making the pots’ led to substandard pots that often broke in the making, and a lack of time for elaboration of designs led to the ‘simplification of shapes and decoration’ evident in recent phase ceramics (Allen 1984:423). It is this emerging but still-uncertain history and its ‘wider social and economic ramifications’ (Allen, 1984:426) that the sites of the Study Area can potentially elaborate, test or rewrite. 4.5.8.4 Pamela Swadling Swadling (1977:38) states that by 1977, about 400 archaeological sites were known from the coastal lowlands of the Central Province by the PNG National Museum and Art Gallery; the oldest of these (e.g. Nebira 4, Eriama 1; susbsequently, Loloata Island) dated to around 2000 years ago, indicating the rarity and great difficulty of finding older cultural materials, despite well-documented archaeological deposits tens of thousands of years old in the Highlands. She further notes that at the time of early European contact, ‘the largest villages were those of the Motu; but from Pari westwards, all Motu villages also had Koita residents … The Koita however had other settlements located on the coastal lowlands inland from the coast, or on hills overlooking the sea’ (Swadling 1977:37). Swadling and Kaiku (1980) excavated two sites near the Study Area: a ‘fireplace in the clay surface of an eroded village site in the Papa salt pans’ (Swadling and Kaiku, 1980:86), dated to 1280±170 BP; and a large archaeological village site at Ava Garau located on a coastal ridge shortly to the northwest of Boera, dated to 1220±95 BP. The Papa site contains red slipped ceramic sherds typical of the earliest phase of known human occupation in the PNG southern lowlands (e.g. Style I of Bulmer 1978; at Nebira 4 Horizon 3 of Allen 1972– see above). At Ava Garau – which Swadling (1977:39) identifies as an ancestral Boera site – ‘pottery was found which shows that both old and new pottery ideas were used by people living there 1,200 years ago. … The influence of new potting ideas, especially in bowl decoration and rim shapes, from the D’Entrecasteaux, Amphlett and Goonenough Islands cannot be denied’. Although the results of archaeological excavations at neither the Papa site nor Ava Garau have been published in any detail; they nevertheless confirm the presence of ancient archaeological deposits capable of shedding considerable light on the history of occupation, ceramic traditions and social interaction for the broader Port Moresby region within the immediate Papa-Boera area. Swadling (1977:42) concluded that while the ancient ceramic assemblages of the broader Port Moresby region showed close formal and decorative affinities with those of the D’Entrecasteaux, Amphlett and Goonenough Islands as well as Milne Bay, Motuan history could not be reduced to recent or foreign arrivals ‘to the shores of Port Moresby’ (as Allen similarly concluded for the last 800 years, see above). Rather, oral traditions ‘do not tell of a far away homeland, but of old village sites along the Central Province coastline. Some of these old villages are said to be very old, whereas others have been recently settled’. It is this history of arrivals, continued or intermittent long-distance contacts, and local practices that lies buried in the archaeological sites of the Port Moresby region. The ceramics at those sites hold particular importance due to the cultural associations they can reveal between sites and regions through time. It is these ceramic conventions that have long formed the focus of archaeological research across the region and beyond. Swadling (1980) has divided the Port Moresby region ceramics into three phases: Early Period (a.k.a. late ‘Red Slip’, ~2000-1200 years ago), Middle Period (a.k.a. ‘Boera-Taurama-Motupore’, ~1200-300 years ago) and Late Period (a.k.a. ‘traditional Motu’, last 300 years). She argues that major stylistic changes in ceramic designs took place between the late Early Period and the Middle Period (broadly B. David, B. Duncan, M. Leavesley 4-520 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project but imprecisely corresponding to the ‘Papuan hiccup’ of Rhoads [1982:146], ‘hiccup’ of Irwin [1991]; ‘ceramic hiccup’ of Summerhayes and Allen [2007]; and ‘hiatus’ of Allen [1997], see above; see section 4.5.8.5 below). Her study of the sources and antiquity of a small sample of the ceramic vessels found in Central and Gulf Provinces archaeological sites (including sherds from Daugo Island site AAQ, the Papa Salt Pan site AWL, and Ava Garau [AMH] near Boera) indicates that ‘early Middle Period sites do not seem to extend as far west as those of the late Early Period. Does this reflect some settlement changes in the Gulf or the impact of the changing situation in the Central Province, as the early Middle Period marks a rather abrupt, but not total, stylistic change in the Port Moresby region’ (Swadling 1980:108-9). She continues (Swadling 1980:115): … the people living at the late Early Period sites in the Port Moresby region were using a number of different clay sources. Why the people living at Ranvetutu were using pots made from Boera clay, rather than clay from near their own village, is not known. … The intricate decoration and complex shapes of the pots made during the late Early Period indicates that considerable time and effort was spent on pot making. These people were certainly not involved in the quick, mass production of pots which occurred in the Port Moresby region at the time of contact. Swadling clearly suggests major cultural change across the Port Moresby region between the late stages of the earliest ceramic phase and the classic Motuan ceramic tradition familiar to us from ethnographic times, changes akin to those questioned by Allen concerning the period between 1200 and 800 years ago in particular (but see section 4.5.8.7 below for a different timing from the Gulf Province) and best investigated through the rich archaeological deposits of the region. Structural remains (postholes) from this Early Period have been found at the Papa Salt Pan site. Furthermore, further to the west in the Gulf Province sites receiving Central Province ceramics, ‘the bulk of the late Early Period potsherds … come from sources in the LeaLea-Boki area. None come from Boera’. Swadling (1980:119-21) continues: The same pattern with most coming from LeaLea-Boki and none from Boera continues in the early Middle Period potsherds from Tei Hill … This finding suggests that the same clay sources continued to be used during the rather abrupt, but not total, ceramic stylistic change which occurred between the late Early Period ceramics in the Port Moresby region. No settlement sites with early Middle Period ware are known from the LeaLea area, but it would not be unrealistic to envisage the continued use of this clay source by descendants of people who may have moved to reside in the Boera village complex from the LeaLea area. … Perhaps the biggest surprise of all, is the lack of late Early Period and Middle Period sherds made from Boera clay in the Gulf sites. … This seems contrary to the widely acknowledged Motuan legend which claims that the hiri was started by Edai Siabo from the Boera area. … The results to hand would indicate that it was the people formerly resident in the LeaLea area, who may have been responsible for producing, using their former clay sources, most of the early Middle Period ware which reached the Gulf. While the people using the Boki clay source in the LeaLea area were the main suppliers to the Gulf of both Early Period and early Middle Period ware, the coming of the Middle Period seems to mark a ?total decline in the movement of Central Province pots to the Gulf. The author is not aware of any middle Middle Period [ware] … having been collected in the Gulf. In other words, it would seem that soon after the founding of the huge village complex at Boera, that potsherds dating to that period no longer appear in the Gulf. A likely explanation is that the oral traditions (including the legendary Edai Siabo story) relate largely, if not entirely, to the most recent phase (last 500 years) of cultural activity in the Gulf and Central Provinces, in line with David’s (2008) suggestion that the history of the PNG lowlands is best understood as a sequence of pulses in occupation and long-distance maritime (ceramic) trade rather than as singular long-term trends. An adequate resolution of this long-standing archaeological conundrum as to the place of Lea Lea and Boera clays and ceramics in long-distance exchange and regional occupation further requires investigation of archaeological sites between Lea Lea and Boera. B. David, B. Duncan, M. Leavesley 4-521 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project A related question that has dogged the archaeology of the southern PNG lowlands concerns whether or not a hiatus in human occupation and long-distance maritime trade occurred around 1000 years ago. Swadling (1976:1) poses this question for the Port Moresby region, pointing out that ‘The excavations and surveys of Bulmer, Allen and Vanderwal along the central south Papuan coast all suggested that there was a chronological break about 1,000 years ago’. A paucity of radiocarbon dates on individual pieces of charcoal (thereby avoiding the potential mixing of charcoal pieces of varied ages) notwithstanding, Swadling (1976:2-3) suggests that the Ava Garau radiocarbon determination near Boera ‘removes the likelihood of a hiatus in the Port Moresby sequence’, and instead ‘suggests continuity into what has been called the Boera-Taurama-Motupore tradition’, as the Boera-Taurama-Motupore tradition is interpreted as a local development of earlier (imported) ceramic manufacturing conventions of the Port Moresby region (in line with Allen’s [1977] interpretations, outlined in section 4.5.7.4 above). Like Allen (1977), Swadling (1976:4) suggests that the BoeraTaurama wares are ancestral to recent Motu ceramics as documented ethnographically. Nevertheless, the question of a hiatus in regional occupation and long-distance ceramic trade between 950-500 years ago remains for the Kikori River area of the Gulf Province. Disruptions in settlement systems, trade relations, and ceramic production in the pottery-producing Port Moresby region villages is key to understanding the lull in ceramics and paucity of known archaeological villages between the occupational pulses in the Gulf Province (see section 4.5.8.5 below). 4.5.8.5 Jim Rhoads, David Frankel and Bruno David: the Gulf Province Sites This brings to the fore a dominant theme of southern lowland archaeological research in PNG, one that is of close relevance to the archaeology of the Central Province and of the Study Area as outlined above: the nature and antiquity of the ethnographically documented hiri trade system as visible from the Gulf Province, recipient end of the hiri trade. As ceramics have been the single-most informative artefact type allowing the tracking of the hiri system’s history, the reporting of ceramic sequences has been of utmost importance for understanding southern PNG’s cultural history. Since the late 1960s when professional archaeological investigations were initiated in southern PNG (e.g. Allen 1972; Bulmer 1971, 1978; Irwin 1985; Vanderwal 1973, 1976, 1978), research has thus focused on understanding ceramic sequences both within the pottery-producing (see Allen 1977a, 1977b, 1978, 1984; Allen and Rye 1982; Bulmer 1982) and receiving (see Frankel et al 1994; Rhoads 1980, 1994) ends of the hiri system. Despite this considerable effort – particularly concentrated through the 1970s into the early 1980s – and significant findings, including the identification of 2000 years of pottery production and trade between the Central Province in the east and the Gulf Province in the west, few excavations and ceramic sequences have been reliably radiocarbon-dated or systematically published, making it difficult to characterise, adequately model, or trace the evolution of ceramic sequences within and between the Gulf and Central Provinces. Initially archaeological researchers who tried to investigate the origins and history of the hiri generally concluded that the hiri system itself (in the form that we have come to know it from ethnography) began only a few hundred years ago, with viewpoints ranging from around 700 to 300 years ago depending on the region at stake, the specific archaeological site, and the kind of evidence used (e.g. oral traditions, archaeological ceramics, archaeological evidence for settlement intensification and population increase). There has, however, also been widespread recognition that the hiri is only one of a number of post-Lapita long-distance Melanesian maritime trade systems operating during the late 1800s (e.g. see Irwin 1985 for discussion of Mailu trade to the east; Harding 1967 for Vitiaz Straits; Uberoi 1962 for the Kula system of the Trobriand Islands), and whose ancestry in the region emerges from at least 2000 years of ceramic history (cf. the sites of Nebira 4, Loloata, Oposisi, Eriava, and perhaps Samoa: Allen 1972; Bulmer 1978; Rhoads 1980; Sullivan and Sassoon 1987; Vanderwal 1973; see Allen and Macintyre 1990 for a review; see McNiven et al 2006 for a critique of the dating of the earliest southern PNG pottery). This extended history of long-distance maritime trade has led to a rethinking of the history of the hiri and a general recognition that across the southern PNG coastline, an early phase of widespread ceramic decorative styles and shapes (termed EPP, or Early Papuan B. David, B. Duncan, M. Leavesley 4-522 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Pottery by Summerhayes and Allen 2007) tends to be separated from a recent phase of highly specialised, regionalised ceramics (and other goods) by a period of ceramic evolution lasting a few hundred years (and dating to various times between about 1200 and 500 years ago, depending on the region; see discussions of Jim Allen’s and Pamela Swadling’s work in section 4.5.8.3 and section 4.5.8.4 above). This latter phase of ceramic transformation was coined the ceramic ‘hiccup’ by Irwin (1991; see also the ‘Papuan hiccup’ of Rhoads 1982:146), and in some regions may have involved abandonment of settlements, a lull in long-distance maritime trade, or a change in occupational systems, such as the hiatus in the cultural sequence at Yule Island/Hall Sound between 1200 and 700 years ago (however the presence of apparent ‘hiatuses’ may simply reflect insufficient, or insufficiently precise, radiocarbon dating; Vanderwal 1973; see also Irwin 1991; Rhoads 1982). In the Gulf Province, Frankel et al (1994), largely following Rhoads (e.g. 1980, 1982:143), identified a hiatus in occupation and/or pottery as a relatively late occurrence compared with the ‘ceramic hiccup’ phase (that is, the period between the end of the EPP and the commencement of the most recent regional ceramic styles) of other sequences further to the east, pointing out that: ‘No sites in the Gulf have been securely dated between 700 and 500/400 years ago. This is probably a product of the limited amount of research and the difficulty of locating sites without pottery, but may well reflect this decline in long-distance trade, at least in pottery’ (Frankel et al 1994:46). Most researchers (e.g. Allen 1977; Swadling 1976) have suggested that the ethnographic hiri immediately post-dates this ‘ceramic hiccup’ phase of 1) transformation in pottery styles (in Central Province pottery-producing communities) or 2) apparent ceramic absence (in Gulf Province pottery recipient communities), and is probably only 500 to 300 years old, including the first direct or indirect arrival of hiri trade pots into the western parts of the Gulf Province (Rhoads and Mackenzie’s [1991] ‘Recent Ceramic’ phase). These and other related historical trends are usually taken as indicating 500-300 years of continuous trade, an increasing standardisation of trade goods (including increasing specialisation and centralisation of ceramic production within the ceramic producing areas), population increases and the establishment of large settlements in the Gulf Province (e.g. Allen 1977a, 1977b; Frankel et al 1994:45-47). Recently, David (2008) has demonstrated major shifts in ceramic trade into the western sections of the Gulf Province 2 beginning 500 years ago , attributed to the onset of the hiri continuing uninterrupted into ethnographic times. This most recent pulse in occupation, ceramics and radiocarbon dates in the Gulf Province, dated to 500-0 years ago, corresponds well with Rhoads and Mackenzie’s (1991) Recent Ceramic and Proto-historic phases. This period of time contains the greatest number of ceramic sherds, traceable to the onset of the ethnographically documented hiri system (again in agreement with Rhoads and Mackenzie’s earlier interpretations). Precisely how the newly excavated ceramics from this most recent period formally, decoratively, economically and occupationally relate to the earlier ceramic phases – in particular how they relate to an earlier pulse of high archaeological representation 1450950 year ago – remains a matter of debate. Archaeological results from the Gulf Province – the receiving end of the hiri trade – have considerable implications for understanding social dynamics across the entire region of the hiri system, including the Port Moresby region. Ethnographically, the annual arrival of the lagatoi fleets was eagerly anticipated in the host villages. Recipient communities expected Motu trade partners to arrive with appreciable numbers of high quality goods, from which redistributions were scheduled with visiting trade partners from hinterland villages. In order not to jeopardize established trade partnerships, the Motu traders themselves targeted those Gulf villages where trade partnerships had previously been established and where subsequent visits had been scheduled by mutual agreement (cf. Oram 1982). The viability of such formalized exchange alliances required the existence of stable settlement locations as destinations for voyaging hiri traders, as bases for negotiation and exchange, and as trade depots through which redistribution could take place. It is at these established centres that the Motu traders arrived, and it is here that villagers amassed the necessary bundles of sago and canoe hulls for 2 ‘Years ago’ and ‘cal BP’ are used interchangeably in this section, although technically this relates more accurately to ‘cal BP’. B. David, B. Duncan, M. Leavesley 4-523 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project exchange with the newly arrived hiri trade cargoes. The mass production of sago starch for exchange by individual trade partners itself necessitated annually scheduled community planning, again requiring the existence of stable settlement locations across the broader region for storage of largescale accumulations. Ethnographically, the hiri lagatoi were partly rebuilt in these trade centres, where Motu traders resided over a period of weeks to months – security of residence and trade locale were essential for successful and sustained trading (e.g. Oram 1982). The sustainability of long-distance maritime trade thus required a significant degree of settlement stability enabling prior planning for the large-scale movement of ceramic wares and sago bundles between distant communities. It is probably no coincidence, therefore, to find that the most recent major pulse in radiocarbon dates in the mid-Kikori River (500-0 years ago) is associated with large quantities of ceramic sherds (of a kind associated with hiri trade wares) in relatively long-lived village locations. Similarly, the only other major pulse in occupation in that region – dated at 1450-950 years ago – is also associated with large quantities of imported ceramic sherds, suggesting the existence at that time also of stable and sustained trade partnerships involving incoming ceramic wares. The implication is that at both these times long-distance trade partnerships were established. Trade partnerships between incoming ceramic traders and coastal Gulf Province villagers, as well as redistribution partnerships between inland and coastal villagers, had to be sufficiently established and reliable for long distance maritime traders to regularly make the dangerous sea journey to the Gulf across distances of 400 km or more. In agreement with Frankel et al (1994), the onset of the hiri trade during the last 500 years or so, as well as the establishment of antecedent trade partnerships between 1450 and 950 years ago, thus implicate more than the trade of material objects. They engendered a broader package of social conditions including the scheduling of community activities, the fostering of inter-regional relations and the establishment of stable village locations as agreed-upon centres of negotiation. With this came articulating village networks within the Gulf, and forms of social and demographic momentum regulating and directing village growth and broader networks of social interaction. While the major pulses in occupation in the mid-Kikori region suggest the existence of such forms of social dynamism 1450-950 and again 500-0 years ago, they also indicate a loosening of village stability presumably in concert with a breakdown in long-distance trade relations between 950-500 years ago, which is so-far characterized by an absence of (imported) ceramics. It is significant to note that this period in the mid-Kikori region lies largely within the ‘ceramic hiccup’ phase of the Central Province – a period of transformation of pottery styles in the ceramic production end of the hiri system. The paucity of radiocarbon dates and the apparent absence of ceramics between 950 and 500 years ago in the Kikori River region may thus reflect contemporaneous and/or shortly earlier disturbances in ceramic producing sites and cultural sequences further to the east. If the precise dating of cultural sequences in the Central Province sites is correct (which is not certain), the rejuvenation of ceramicsago exchange in the Gulf Province around 500 years ago appears to post-date the start of intensified pottery production and the most recent ceramic phase (immediately following the ‘ceramic hiccup’) in the Port Moresby area by perhaps 200 years (there possibly involving Koita-Motu displacements; Allen 1977a; Bulmer 1978). During this most recent period, the establishment of a new phase of trade partnerships and stable settlement locations were associated with new forms of regionalized ceramics, indicating a break-down of the earlier and more widespread ceramic conventions. Critical to understanding the onset of this new phase is, therefore, the period we have come to know as the ‘ceramic hiccup’, for it is this period of transformation that represents the link between the earlier and later phases of ceramic production and long-distance maritime trade. In such ways the archaeology of distant (Gulf Province) places has profound significance for understanding the socio-cultural history of the ceramic-producing villages in the Port Moresby region, and vice versa. The archaeological significance of the hiri trade and its historical antecedents has recently been reframed and set in new focus by the findings of red-slipped ceramics in northern Australian waters B. David, B. Duncan, M. Leavesley 4-524 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project (Torres Strait). At Ormi and Mask Cave, Carter et al (2004) and McNiven et al (2006) have found stratified ceramic sherds on islands that have no ethnographically known pottery making (or using) traditions. The significance of these findings is highlighted by McNiven et al’s (2006) claims for the presence of ceramic sherds dated to 2400-2500 years ago from Mask Cave on the islet of Pulu, which they suggest may relate to the onset of southern PNG influences into Torres Strait around 2600 years ago. Sourcing of the sand tempers by Dickinson (in McNiven et al 2006) failed to specifically locate the manufacturing centre(s), but was tentatively identified to western Torres Strait sandy-clay sources. If the dating at this site is correct, the Mask Cave results would pre-date any confirmed ceramics along the PNG southern coast, and likely indicate the presence of Austronesian maritime colonizers akin to the early Motuans traveling as far west as Torres Strait. One problem in the resolution of the origins and history of ceramics in southern PNG and northeastern Australia remains the inadequately dated antiquity of human occupation and cultural sequences along the entire southern coastline of PNG, the Port Moresby area included (see McNiven et al 2006). The temporal pattern in settlement and ceramics from the Gulf Province is of considerable significance for understanding the region’s social history and its connectedness with ceramic production centres in the Central Province area, including Lea Lea-Boera. What we can say from the occupational trends in the Gulf Province is that settlement systems were never stable for very long: occupational trends are best understood in terms of pulses rather than long-term trends, and these pulses appear to be associated with the operation of long-distance Motuan trade networks coming from the broader Port Moresby area. Because of the workings of the hiri system, cultural sequences in one part of southern coastal PNG are closely linked to those of other parts, even if many hundreds of kilometres apart (as recognized by previous researchers). In light of these findings, it is likely that ethnographic stories of population movements, village and clan origins for this region relate to the latest (last ~500 years), rather than earlier, phases of occupation or use. What this ethnography also enables us to conclude is that to understand land use in the Gulf Province, we need to know more than environmental conditions and environmental histories, requiring a focus on the specifics of social interactions which, in this case, have come to guide settlement processes; understanding the cultural history of places requires consideration of past social relationships. What the above results highlight is the significance of ceramic producing centres for understanding the history not just of those locations for themselves, but for understanding the history of the entire southern coastal region of PNG, as an inter-connected social network. For this reason, the archaeological sites of the Lea Lea-Boera area, as a renowned domestic and hiri pottery manufacturing centre, attain an extra-regional significance that requires detailed documentation and considered management. We return to this point in section 4.5.8.6 below. 4.5.8.6 Geoff Irwin Geoff Irwin did not work in the Port Moresby region, but his archaeological research in the Mailu area along the coast 260 km to the east revealed historical trends applicable also to the Study Area. Like other archaeologists working along the southern coast of PNG, his basic premise was that ‘One can identify settlement patterns simply by plotting the distribution of archaeological sites shown to be highly similar in their ceramic inventories’ (Irwin 1978:301). Irwin (1978) argued that the history of the Mailu area, as indicated by archaeological research, can be divided into three major periods, which he called Early, Mayri and Mailu. The Early period (2000 to around 1500 years ago) was characterized by a series of pottery-producing villages along the mainland coast and on offshore islands. ‘Through time’ 3 – i.e. during the Mayri period (from around 1500 to 400 years ago ) into the early Mailu period (after approximately 400 years ago, at the time of writing [1978] identified as ‘350 b.p.’ by Irwin) – writes Irwin (1978:299), ‘the density of mainland settlement increased and there was an associated shift in 3 There is some ambiguity as to the timing of the Mayri and Mailu periods, for Irwin (1978:302) also writes that the Mayri period ‘dates some 6-800 b.p.’; that is, that it continues to around 600 to 800 years ago. As Irwin here discusses settlement patterns specifically, it is likely that he is referring here only to the distinctive (regionalised) Mayri settlement patterns lasting until 600 to 800 years ago, rather than to the Mayri period of ceramic conventions (which lasts until about 400 years ago). B. David, B. Duncan, M. Leavesley 4-525 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project village site location. In addition, one settlement began to differentiate from others at a rate which accelerated through time. By the period of European contact, the small island of Mailu was the location of a settlement that can be described as a central place. It was larger, socially more stratified, more influential and functionally specialized than any other place’. During the Mayri period, ‘pottery making was a widespread skill and occurred in several villages’ (Irwin 1978:300). By the time of the early European contact period, the entire region was dominated by a single pottery-making village (on the island of Mailu) holding a monopoly over production and ceramic trade as well as use of large oceangoing canoes, despite the fact that by that time there were many more villages than previously along the coast, and that these villages were more closely but less regularly spaced than during earlier periods (being on average 7, 6 and 3 km apart during the Early, Mayri and Mailu periods respectively) (Irwin 1978:304, 305). Along with this increasing centralization and specialization of ceramic production and trade, and increasing populations and village density also came a move from coastal village locations to hilltops for purposes of defense, a further indication that social relations were significantly different between the latest (ethnographic) phase and earlier times. Because of insufficient radiocarbon dating, Irwin (1978:315) concludes that ‘The major change in pattern occurred between early in the Mayri Period and 1890’ – a period covering from around 1500 to 150 years ago. Clearly and despite significant archaeological results, the archaeological record still holds much to reveal about the region’s past. This process of socio-political development was analogous to that documented among the Motu (and Koita) from the Port Moresby region, whereby archaeologists such as Bulmer, Swadling and Allen have argued for a three-fold (sometimes reduced to a two-fold) developmental sequence from an early phase of widespread ceramic conventions that changed in tandem between regions, indicating ongoing inter-village communication and interactions, into a middle phase of ceramic transformations, leading to a recent (ethnographic) phase of peak populations making regionalized but highly standardized ceramics in poorly articulating villages among the Motu. It is the archaeological records (sometimes combined with oral histories) of each region that is best capable of discovering this deep history, and here it is worth noting that while some (but limited) archaeological research at village sites has been undertaken in the Port Moresby region, as is the case also in the Mailu district ‘little is known archaeologically of the many sites which have been designated as hamlets’ (Irwin 1978:304). Yet a region’s history does not just consist of the villages, but also the other places of human activity, and the smaller (e.g. hamlet) sites are just as important in deciphering a region’s and peoples’ history. The ceramic sherds contained within these smaller settlements may well shed significant light on the history of inter-regional interaction within and beyond the Port Moresby region (taking into consideration also the fact that early Koita settlements tended to be hamlet-sized), as well as the whereabouts of ceramic-manufacturing centres at different times in the past (see also section 4.5.8.7 below). 4.5.8.7 Summary: The Study Area’s Indigenous Cultural Sites as Identified from Previous Archaeological Investigations Bulmer (1978:5) explicitly states in her major study of archaeological sites in the broader Port Moresby area that archaeological sites are widespread and often structurally complex: ‘nearly every hilltop in the region has at least a small scatter of pot-sherds and flakes’, and ‘many sites will have to await accidental discovery if they are buried deeply beneath layers of sediment’. Because ‘indirect trade was carried out all along the Papuan coast … and other trade occurred over the top of the mountains’, while direct trade connected the Central Province with the Gulf Province, she also points out that archaeological sites further to the west in the Gulf Province will be of relevance to those of the Port Moresby region, and vice versa, as per previous discussions in section 4.5.8.2 above (Bulmer 1978:67). Golson (1968:70) similarly concludes that ‘The archaeological information to be collected from village sites … will help to build up a picture of the Port Moresby coast in the centuries before the Europeans arrived’. The sequence of ceramic styles that we find in these archaeological sites should enable us ‘to trace the history of Motu relationships with other peoples inland and along the coast’ (Golson 1968:71). The significance of pottery-bearing archaeological sites in the Port Moresby region has been stressed by all archaeologists who have worked in the region. B. David, B. Duncan, M. Leavesley 4-526 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project The sites that have been archaeologically excavated from the Port Moresby region thus reveal important details as to the kinds of cultural deposits we can expect to occur in the Study Area. Only eight sites have been professionally archaeologically excavated and reported in the Port Moresby area westward to Papa: ACL (Nebira 4) by Jim Allen (1972) in 1969-1970; ACK (Motupore) mainly by Jim Allen, but also by or with Wal Ambrose, Sandra Bowdler, Mary-Jane Mountain, Pamela Swadling, Allen Thorne, Les Groube and John Burton (Allen 1976b; Swadling 1997:2; see also Golson 1968) in 1970 to 1986; AMH (Ava Garau, near Boera) by Pamela Swadling (Swadling and Kaiku 1980) in 1976; AWL (Papa Salt Pan) by Pamela Swadling (Swadling and Kaiku 1980); ACJ (Nebira 2), ACV (Eriama 1) and AJA (Taurama) by Susan Bulmer (1978; see also Lampert 1968); and ANT (Loloata Island) by Sullivan and Sassoon (1987) in 1985. The PNG National Museum and Art Gallery site survey record form for midden site AADI at Boera village indicates that Les Groube undertook an excavation and obtained a radiocarbon date there in 1988, but the results have never been published. Similarly, the site survey record form for the creek-bank pottery site of ARQ shortly to the east of Boera indicates that a radiocarbon date has been obtained from this site by Colin Pain in 1979, but it has not been published (see Table 4.5.6 for site details). Most of these sites had been targeted for research because oral traditions identified them as dating to ancient times; no sites expected to contain exclusively ‘recent’ deposits (and therefore targeting the last few hundred years of cultural history) have been excavated. The oldest deposits have been dated to around 2000 years ago at Nebira 4, Eriama 1 and Loloata Island (see Table 4.5.5). Nebira 2, Taurama, Ava Garau, and Motupore are in themselves each also very old village sites close too, or in some cases more than 1000 years old. All reported excavations contain rich pottery and shell deposits, as well as numerous human burials (at Taurama, no burials were found in the archaeological excavations, but ‘Unrecorded human bones have eroded from the beach front from time to time’ (Bulmer 1978:264). At Nebira 2, more than 55,000 pottery sherds were excavated, along with the remains of at least 45 individuals, although the true numbers present at this site are likely to be around four times these amounts as only about one quarter of the site has been excavated (Bulmer 1978:135). Taurama is a beachside ‘foundation village of the western Motu’ and is said to have been settled from Motupore around 14 generations before 1978 (corresponding well with the timing of abandonment at Motupore as evidenced by archaeological investigations) (Bulmer 1978:258, after Oram 1969:429; see also Golson 1968:69). Taurama itself is said to have been abandoned as a result of warfare with the people of Loloata Island some seven generations ago, resulting in migrations to Manugava (ABU) onwards to Tauata and Badihagwa, ‘the forerunners of the present villages of Pari and Hanuabada’ (Bulmer 1978:258-59). Taurama contains a rich assortment of shells, stone and shell artefacts (including imported obsidian flakes), beads, vertebrate faunal remains, almost 25,000 pottery sherds, and structural evidence (e.g. postholes). At Eriama 1, 48-50 burials were excavated, along with 1530 pottery sherds, shell and animal bone remains, and stone artefacts including a small amount of exotic obsidian, probably imported from Fergusson Island (Bulmer 1978:202, 246). Many of these interments contain burial goods such as shell arm rings, beads, pottery, stone artefacts, or bone or tooth ornaments (e.g. Bulmer 1978:182, 226-34, table 6.9). At Motupore, ‘several hundreds of human skeletons’ have been reported (Lampert 1968:74). Archaeological excavations here have revealed extremely rich cultural deposits (including a specialized bead-manufacturing industry) in addition to skeletal remains. The implication of these archaeological excavations together with ethnographic information is that village sites across the broader Port Moresby region including Lea Lea-Boera, will almost certainly contain human remains from burial practices beneath or near the location of past houses. Rockshelters may also contain human skeletal remains. Other, smaller cultural sites have not yet been excavated from the broader region, so their site contents remain a mystery. B. David, B. Duncan, M. Leavesley 4-527 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.5 Site Nebira 2 Eriama 1 Taurama Nebira 4 Motupore** List of reported excavated archaeological sites in the broader Port Moresby region, with radiocarbon dates relating to cultural activity PNG National Museum and Art Gallery Site Code ACJ ACV AJA ACL AAK Radiocarbon Date Laboratory # References 0±270 (GaK-2674) Bulmer 1978 280±80 (GaK-2672) Bulmer 1978 380±120 (GaK-2675) Bulmer 1978 390±90 (GaK-2345) Bulmer 1978 660±150 (GaK-2673) Bulmer 1978 720±80 (GaK-2346) Bulmer 1978 210±70 (GaK-2671) Bulmer 1978 380±120 (GaK-2668) Bulmer 1978 600±125 (GX-3334) Bulmer 1978 1930±230 (GaK-2670) Bulmer 1978 560±85 (I-6862) Bulmer 1978 775±85 (I-6887B) Bulmer 1978 865±140 (I-6863) Bulmer 1978 880±250 (GaK-2667) Allen 1972 1760±90 (I-5796) Allen 1972 3340±160* (GaK-2990) Allen 1972 330±55 (ANU-1177) Allen 1977 390±65 (ANU-1212) Allen 1977 740±105 (I-5903) Allen 1977 810±80 (ANU-1211) Allen 1977 1010±80 (ANU-1219) Swadling 1997 Loloata Island ANT 2300±100 (shell date) (ANU-4808) Sullivan and Sassoon 1987 Papa Salt Pan AWL 1280±170 (SUA-1524) Swadling and Kaiku 1980 Ava Garau AMH 1220±95 (SUA-515) Swadling and Kaiku 1980 Radiocarbon dates use the Libby half-life of 5570 years (after Bulmer 1978:table 6.3). *Allen (1972:120) rejects this determination in favour of I-5796. **Swadling (1997:2) states that 22 radiocarbon determinations have been obtained from Motupore, but these do not appear to have all been published. Of note here is that Bulmer (1978:345-46) has written that, based on the distribution of surface and/or excavated ceramics, there is evidence of a substantial village at Boera during what she calls the Middle Period (about A.D. 1000-1500) and at Lea Lea during the Proto-historic period (about A.D. 1500-1875), although little archaeological research (and no reported excavations except for Ava Garau and Papa Salt Pan) has been undertaken in the Port Moresby region northwest of Boera and therefore the deeper (older) cultural deposits of this area in particular remain largely unknown. In summary, archaeological site patterning for the broader Port Moresby region indicates the following for the general area including the Study Area: • A range of site types can be expected, including village sites; clay, ochre and stone artefact sources; freshwater wells; hunting camps; gardens. B. David, B. Duncan, M. Leavesley 4-528 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project • Village sites can be expected along or near the coast (due to the maritime orientations of the Motu) as well as on hilltops (due to the defensive positioning of ancient Koita settlements). The former in particular will almost certainly be stilt villages, possibly with the presence of dubu. • Large village sites are likely to occur on the coast, with smaller hamlet-type settlements further inland. However, migrating coastlines through time may mean that at some times in the past coastal villages were located further inland than today’s coastline would indicate. • The inter-tidal zone and near-shore environment is likely to contain the remains of past settlements. • ‘Prehistoric communities exploited not only the coastal zone and its range of resources, but also the resources of the inland river plains’ (Bulmer 1978:326). Garden sites are likely to occur in the hinterland. • Archaeological sites could be buried deeply below the present ground surface (due to alleviation subsequent to site occupation, partly a result of siltation caused by forest clearance due to hinterland burning and gardening practices). • Burials are likely to be found within any village site, as well as in some caves and rockshelters. • Customary spiritual/story places. • Most of the archaeological sites are likely to occur on land or in the inter-tidal zone, but lagatoi or canoe wrecks and spiritual/story places are present underwater in the general region. Impact considerations and management measures will need to take the above general principles into consideration. Furthermore, we note that no maritime archaeological recording work has previously been undertaken in the Study Area, with the exception of sporadic recordings of WWII crashed plane sites and military sites by amateur enthusiasts and the recording of one aircraft by the Department of Modern History. Before discussing the results of our own field surveys in the Study Area, in section 4.5.9 below we review the literature and archival records for evidence of European contact period, and shallow marine, cultural heritage sites for the Study Area and its vicinity; a list of indigenous cultural heritage sites identified above from the literature review is presented in section .4.5.9.5. B. David, B. Duncan, M. Leavesley 4-529 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.9 Archaeology: European Contact Period Cultural Heritage Description 4.5.9.1 European Exploration Here a very brief outline of European exploration is presented to set the scene as to what kinds of early European contact situation, and archaeological objects relating to this period, may be found in the Study Area. The Portuguese explorer De Meneses first sighted the island of New Guinea along the Bird’s Head coastline (western New Guinea) and landing on Biak Island in 1526. But it was not until 1545 that Spanish explorer De Retes claimed the northeastern shores of the country and named the area ‘New Guinea’, due to the resemblance of the local people to those in the African country of Guinea (Kanasa 2006:10). The Portuguese and Spanish explorers Torres and de Prado landed at nd Red Scar Head on 2 September 1606 in search of fabled gold deposits (Kanasa 2006:11; Whittaker 1971:628). Dutch explorations of the southern New Guinea coast began in 1605 when Willem Janszoon was sent by the Dutch East India Company to investigate the area. He proceeded as far as the entrance to Torres Strait and mistook numerous islands of the New Guinea coastline to be connected to the legendary Terra Australis. Later voyages by other Dutch navigators, including Abel Tasman in 1643, added further knowledge of the southern coastline (Darby 1945:121-22). th Dutch investigations of the island of New Guinea in the 17 century failed to reveal anything considered of worth, while the English buccaneer William Dampier had also surveyed sections of the northeastern coastline in 1700 (Whittaker 1971:629). In 1770, James Cook conclusively proved the existence of a strait between Australian and New Guinea, confirming Torres and de Prado’s 1606 discovery (Darby 1945:1212). In 1788, the French explorer La Pérouse in the Astrolabe planned a voyage which was to include surveys of the Louisiade Islands (previously discovered by Bougainville) and a new route between New Holland (Australia) and New Guinea (Horner 1996:3). However, it is probable that this never took place as his ships were wrecked in the Solomon Islands (Stanbury and Green 2004). Further French expeditions began surveying the New Guinea coastline and interior, and include D’Entrecasteaux (1792), Duperrey (1823-24) and d’Urville (1827-28) (Darby 1945:122). th Early in the 19 century, growing trade with China opened up more sea routes between Sydney and Canton and more frequent contacts between indigenous peoples of PNG and Europeans. New Guinea coastal settlements could supply food and water to ships en route, and it is probable that whalers frequented these shores in search of provisions and fresh crew (Whittaker 1971:630). Although the French (d’Urville in the Astrolabe and Zélée) had surveyed the southern PNG coastline in 1839 (during which they named Mt Astrolabe), they halted their surveys just east of (the current) Port Moresby and did not proceed further west (Lubbock 1968:253-55). In 1846 the British Navy planned further surveys of the southern New Guinea coast to continue earlier explorations undertaken by Captain Blackwood of the HMS Fly, who had documented the Fly River estuary along with the western shore of the Gulf of Papua in 1842. A survey was thus planned to extend from the Louisiade Archipelago (in the southeast of PNG) to Cape Valesche (Lubbock 1968:163; Darby 1945:122). The influence of the Dutch East Indies Company, which was guarding its monopoly in the Spice Islands, had effectively prevented most previous attempts to survey the PNG southern coast. In 1849, Captain Owen Stanley in the HMS Rattlesnake, along with Captain Yule of the HMS Bramble, undertook a survey of the southern coastline of New Guinea (Figure 4.5.19). During this time Captain Stanley named a high mountain range after himself (Owen Stanley Ranges), and the highest peak after his predecessor in the area, the French explorer d’Urville (this peak was subsequently renamed Mt Owen Stanley by hydrographer Sir Francis Beaufort in MacGillivray’s 1852 map, but was again renamed as Mt Victoria in 1888). Stanley named Redscar Bay and Head after Preston in the U.K. and B. David, B. Duncan, M. Leavesley 4-530 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project was impressed by its anchorage. Stanley had avoided the shallow shoal waters of Caution Bay and most of the coastline due to fears of being shipwrecked or massacred by the local Indigenous peoples (Figure 4.5.20). His survey from Rossel Island to Cape Possession was released in 1850 by the Hydrographic Department of the British Admiralty (Lubbock 1968:242, 253-56). In 1873, British Naval Captain John Moresby undertook a survey of the southeastern and northeastern sections of the PNG coast, predominantly for military purposes (Kanasa 2006:10; Darby 1945:122). Local villagers from Lea Lea provided the story behind the naming of Caution Bay by Europeans in the th late 19 century. The name relates to the practice by coastal traders of extinguishing all ship lights when sailing through the area at night for fear of Lea Lea villagers rowing out and killing all aboard (Igo Meauri, personal communication 2008). B. David, B. Duncan, M. Leavesley 4-531 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.19 Huxley's Map of the New Guinea Coast drawn from surveys by HMS Rattlesnake in 1849 (Lubbock 1968:211) B. David, B. Duncan, M. Leavesley 4-532 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.20 MacGillivray's 1852 map of the south coast of Papua New Guinea, showing Caution Bay area in the red inset (Lubbock 1968:248) As noted, prior to 1870 occasional vessels involved in the Australian trade or whaling would occasionally visit the southern New Guinea coast to obtain wood, water and other supplies. However, from 1865-1870 a regular trade network began to be established as pearling and bêche-de-mer (sea cucumber) camps were established in the Torres Strait islands. By 1870, Australian vessels began visiting the southeastern New Guinea mainland to trade for cedar timber, which was profitable until at least 1880 (Darby 1945:123). A commercial trading station and plantation established on New Britain in 1873 by English interests established a benchmark that led to the introduction of widespread trade and settlement across New Guinea. Further interest was expressed in the region in the 1870s by German, English and Australian commercial speculators as potential settlements for agricultural and mining developments. Although the discovery of gold near Port Moresby in 1877 prompted a rush of miners to the area, fever, hunger and difficult conditions led to abandonment of the search. During this time the Queensland government provided an agent to maintain law and order (Darby 1945:123-25). Missionaries established a presence in the Port Moresby area by 1874 as a direct result of Moresby’s earlier surveys, when Reverend Lawes became the first European from the London Missionary Society to be stationed in New Guinea. The missionaries travelled around the local coastal villages learning the languages and gaining their confidence as a prelude to establishing permanent local missionary teachers in the region. By 1883, New Guinea was involved in the native labour trade, which saw many local people transported offshore to work on plantations around the Pacific (Darby 1945:123-25, 233). By 1875, the significance of New Guinea as a strategic asset for the defence of Australia was recognised by the British Government. With the rising interest of other colonial powers in the Pacific (France, Russia, Germany and the United States) and fears of a Russian invasion associated with war in the Crimea, New Guinea was established as a British Protectorate in 1884 under the authority of Sir Peter Scratchley (who had designed Australia’s defence systems). During this period Scratchley controlled shipping and introduced laws to bring order between the European and local populations. The region became an Australian Territory in 1906 (Kanasa 2006:21; Darby 1945:125-29). B. David, B. Duncan, M. Leavesley 4-533 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project In 1945, the southern coastline of PNG still required extensive surveying to provide adequate information for mariners. The principle exports of Port Moresby at this time were associated with livestock, copra, and timber, with smaller outward cargoes of desiccated coconut, coffee and rubber. Regular traders visited Port Moresby (including those of the Burns Philp Company), maintaining services to Australia on a three-weekly rota, along with other companies who provided services to the Philippines, East Indies, Japan, China and New Zealand (Darby 1945:123, 235-36, 245). 4.5.9.2 Missionary Activity In 1870, Reverend John Williams of the London Missionary Society (LMS) issued directives for Reverend Samuel MacFarlane to establish a presence in New Guinea. On 20 May 1870, an LMS party sailed from New Caledonia. In 1871, Reverends MacFarlane and Murray, along with eight New Caledonian teachers, established missions in the islands of Torres Strait. In 1872, Reverend Murray landed at Vari Vari Island (near Redscar Head) and was met by local people in canoes. After some initial reluctance, they came aboard the LMS schooner and good relations were established. The missionaries then set about reconnoitring the area with a view to establishing the site for a new mission. After initially visiting Kido but finding it unsatisfactory for the proposed mission, they established a small mission at Manumanu in 1872, where the first missionaries to reside and teach in PNG were settled. Reverend Lawes became the first European to live with Papuans on the mainland when he established a further station at Elevala in 1874 (Kanasa 2006:18-20). The Reverend James Chalmers then landed at Boera on 12 October 1877 (Stewart 1973:277). By 1890, the administrator of British New Guinea William MacGregor had outlined a policy for all religious denominations in British New Guinea to avoid conflict with each other. The ‘sphere of influence’ policy allocated the whole of the southern PNG mainland except Yule Island and Mekeo district, inland as far as Boku, to the LMS. By 1909, the LMS had established an influential presence in the area, with attendance numbers at the mission schools rising to 5000 (Kanasa 2006:18-20). There is to this day a strong Christian influence in all three villages surrounding the Study Area. 4.5.9.3 Sisal Farm/Lea Lea Airstrip/Fairfax Cattle Station After the annexation of New Guinea by Australia from the British in 1906, large tracts of land were acquired for plantation purposes. By 1909, it was recorded that there were over 139 plantations in the Australian territory growing coconuts, rubber, sisal-hemp and coffee (Kanasa 2006:25). The landward area of the Portion 152 area was formerly a sisal plantation which produced hemp. Beginning in 1913, the farm used a light railway to transport the harvested crop. The venture lasted until 1923 when the high cost of shipping and a drop in sisal prices led to its decline (Stuart 1973:277; Douglas 2007:54). An airstrip was built in this location prior to WWII to service the plantation, and was used by Motuan businessmen to fly light planes to Kokoda and Yoda (Ben Moide, personal communication 2008). The area was later used as an emergency airstrip (known as Lea Lea airfield or Schimmer Airstrip) during WWII, from at least 1942 onwards (Ben Moide, personal communication 2008; Douglas 2007:54; see DoMH display, 2008). At the end of WWII, the area was converted to a cattle farm run by the Fairfax family (Stuart 1973:277) which ran until the 1980s (Douglas 2007:54). 4.5.9.4 Military and Defence History Papua and New Guinea were pivotal strategic and tactical locations for the battle of the Pacific during WWII. An enormous build-up of defence armoury and other mobile military assets were stationed in this region during the conflict. A more extensive history of the role of the Port Moresby region in WWII is required to give an overview of the range of activities that took place in this area (see below). PNG and the Second World War (by Bob Marmion, military historian, University of Melbourne). Papua New Guinea has long attracted foreign interest because of its natural riches and strategic advantages. As a result, from the 1870s until Independence in 1975, the island has been occupied by a series of foreign powers including Britain, Germany, Australia, Japan, and again Australia. B. David, B. Duncan, M. Leavesley 4-534 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project The late 1800s was an era when the Australian colonies feared European expansion into the South Pacific region; therefore in 1883, the Queensland government made an abortive attempt to annex the eastern part of New Guinea to prevent further German expansion southwards (Meaney 1976:17). In 1884, the German government formally took possession of New Guinea, ostensibly to develop copra. Though the annexation by Queensland the previous year was declared void by the British government, Britain later (1888) annexed the southern half of eastern New Guinea outright. By 1902, the protectorate was placed under the control of the new Commonwealth of Australia. In 1906, this area became the Territory of Papua (Brune 2004:10). In 1914, at the commencement of the First World War, the German colony of New Guinea was occupied by Australian troops. Following Germany’s defeat, Australia was given a mandate by the League of Nations to govern the former German colony (Brune 2004:10). This mandate remained in place until the Japanese invasion during the Second World War, when civilian administration of Papua New Guinea was replaced by Australian military control based in Port Moresby. In 1945, civilian administration recommenced under the Australian government and continued until Independence was granted in 1975. The Pacific War 1941-1945 (by Bob Marmion). Concerns over possible Japanese expansion into th southeastern Asia dated back to the turn of the 20 century (Walker 1999:75, 230). However, as a result of a series of non-aggression and trade treaties and the fact that Japan fought during the First World War on the Allied side, European powers during the 1920s and early 1930s tended to underestimate the rising Japanese militarism and the potential for conflict in the South Pacific (Walker 1999:75, 230; Meaney 1976:167). While Japan was seen as a threat, as evidenced by the establishment of a major British naval base at Singapore, European focus was squarely on fascism in Germany and Italy (Robertson 1988:238-39). By the late 1930s the European powers, along with Britain, the United States and Japan had commenced rearming in expectation of the outbreak of another World War. In 1937, Japan commenced an eight year war with China. A year after the outbreak of WWII in September 1939, Japan signed the Tripartite Pact, thereby allying itself with Nazi Germany and Italy and potentially setting on a collision course with Britain, the United States and the Dutch Government in exile (which still controlled the resource-rich Netherlands East Indies). In July 1941, 50,000 Japanese troops were moved into French Indo-China as a precursor to any attack on the Dutch East Indies. During the course of 1941, relations between Japan and the Allies deteriorated. In an attempt to curb Japanese militarism, the Allies severely restricted the exportation of fuel and other resources vital to the Japanese economy. As Japan imported most of her industrial raw materials such as oil, iron ore and steel, this was seen by the Japanese as a direct attack on their economy. To the Japanese the choice was simple: face immediate economic collapse or withdraw its troops from China and Southeast Asia with its attendant loss of face. Such a withdrawal, in the face of a continued Allied embargo or slowing of raw materials, would have the same effect in the long term. The Japanese therefore commenced planning for a war in the Pacific against the British and Americans (Figure 4.5.21). On 7 December 1941, Japan attacked Pearl Harbour, thus bringing the United States into the war. Hostilities between Japan on one hand, and Britain and the Dutch on the other, followed shortly afterwards as the Japanese attacked Malaysia, Hong Kong, Thailand, the Dutch East Indies, New Guinea, Burma and the Philippines. In February 1942, following the Malay campaign, Singapore fell. By the middle of April, Japanese forces occupied much of Burma, Borneo and the Dutch East Indies including western New Guinea. The Australian mainland came under regular air attack. In May 1942 the Philippines surrendered. In effect, as a result of their rapid conquest of Southeast Asia, the Japanese had created a buffer zone that not only provided the natural resources required by Japanese industry, but also a defence in B. David, B. Duncan, M. Leavesley 4-535 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project depth. By attacking Pearl Harbour, the Japanese had intended to destroy the American Pacific fleet and thereby allow the Japanese to over-run those countries which could supply it with the natural resources it required. Rather than trying to defeat the Americans, initially Japan hoped to achieve a negotiated peace. Knowing full well that there would be retaliation and that Australia would be the most likely staging post for any counter-attack, Japan attempted to isolate the Australian mainland from both Britain and the United States (Milner 1957:12). By occupying the Dutch East Indies and New Guinea, the Japanese were well positioned to intercept any Allied troop and material build up in Australia as a precursor to a counter-attack. Figure 4.5.21 The Japanese objectives 1941-1942 (http://www.awm.gov.au/underattack/index.asp) Having captured a number of important positions which provided harbours and airfields (such as Rabaul, Lae and Salamaua) by March 1942, the Japanese ideally needed to occupy Port Moresby to ensure the isolation of Australia. Port Moresby would also provide a valuable link between the Japanese forces in the Dutch East Indies and the islands to the north and east of New Guinea. During WWII, Port Moresby was the key port and administrative centre in Papua (Robertson 1984:135). Already a major trade centre for local markets when Papua was annexed by the British, Port Moresby slowly developed over the years as port and administrative facilities were established. Following the outbreak of the Pacific War in December 1941, Port Moresby underwent a dramatic change as it became the nerve centre for Allied operations in Papua. There was a significant build up of Allied troops so that by early 1942 the city had become the base for thousands of troops and airmen. Air attacks could be mounted against Japanese positions along the north coast of Papua. Port Moresby therefore became a major target. The first attempt to capture Port Moresby occurred in April/May 1942, when the Japanese despatched a fleet including two aircraft carriers and troops transports, under the command of Admiral Takagi, to attack both Port Moresby and the Solomon Islands (Figure 4.5.22). The Japanese fleet was th intercepted in the Coral Sea and defeated on 7-8 May 1942. A month later at the Battle of Midway, the Japanese suffered a major defeat which, as a result of losing a number of aircraft carriers, B. David, B. Duncan, M. Leavesley 4-536 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project severely curtailed their ability to mount amphibious operations in the South Pacific. The lack of airpower was to play a major role in the Japanese defeat in Papua New Guinea. Figure 4.5.22 Extent of Japanese conquests in 1942 (Holmes 1988) Despite the losses incurred at Midway, a second attempt to take Port Moresby occurred in July 1942 when Japanese forces were landed on the north coast of Papua at Buna, Gona and Sanananda. The resulting campaign between July 1942 and January 1943 saw a series of savage battles over the Owen Stanley Ranges. Initially the Japanese, under the command of Major General Horii, managed to fight their way along the Kokoda Track to within sight of the lights of Port Moresby before being halted by Australian troops. The Australians opposing them had originally consisted of militiamen, only partly trained, inexperienced and with little in the way of heavy weapons. These Australian militiamen mounted a slow fighting retreat. Eventually, reinforced by seasoned soldiers of the Australian Imperial Force, the Japanese were halted. With the Japanese supply lines overstretched, disease-ridden, th starving and with no hope of reinforcements due to the fighting on Guadalcanal, on the 26 September 1942, Horii had no choice but to obey an order to withdraw back along the Track (Steiberg 1978:51) (Figure 4.5.23). B. David, B. Duncan, M. Leavesley 4-537 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.23 The Kokoda Campaign 1942 (Steiberg 1978) th Japanese problems were further compounded by a defeat at Milne Bay on the 15 August 1942. Over the course of two weeks, the Japanese suffered their first defeat on land at the hands of the Australians (Steiberg 1978:50). Milne Bay also marked the southernmost advance of Japanese forces during the War. In November 1942, the Allies attacked the Japanese beachheads at Buna, Gona and Sanananda in what came to be known as the Battle of Buna-Gona. By January 1943, the Japanese forces in Papua had been eliminated, thus removing the threat against Port Moresby. Fighting continued, however, in New Guinea until 1945. In June 1943, having secured the northern Papuan coastline, the Allies launched Operation Cartwheel. This strategy was designed to isolate pockets of major Japanese resistance, to liberate Rabaul and to enable the island-hopping campaign of the later war-years. The Importance of Port Moresby during the Pacific War (by Bob Marmion). Port Moresby had been the civil administrative centre of the Territory of Papua prior to the Second World War. It provided a direct link between Papua and Australia. By 1942, as General McArthur’s logistical base, Port Moresby had become the most important Allied port in the southern Pacific region (Figure 4.5.24). Located on the southern side of Papua along the main shipping channel between the Indian and Pacific Oceans, it played a vital role as the communications, supply and operational base for Allied land and air opposition to the Japanese thrust south. In addition to Allied command in Papua New Guinea, rear echelon units such as military hospitals, ordnance depots, training wings, communication centres providing links with Australia, transport and a host of other ancillary services were based in Port Moresby. With the development of infrastructure such as electricity, running water, airfields and the construction of a wharf on Tatana Island, the amount of supplies coming through the port had doubled B. David, B. Duncan, M. Leavesley 4-538 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project by October 1942. This in turn dramatically improved the level of air support, reinforcements and supply to forward troops. The development of Port Moresby into a major supply base was a significant achievement. As Samuel Milner noted: The reinforcement of Port Moresby in 1942 was no easy matter. Its supply line from Australia across the Gulf of Papua was exposed to enemy action. Its port facilities were inadequate; its two existing airfields were small and poorly built; and, except for one field at Horn Island in Torres Strait, there were no intermediate air bases between it and the concentration area in the Townsville-Cloncurry region, 700 miles away. After a thorough reconnaissance of Port Moresby, General Casey, General MacArthur's engineer officer, began drawing plans for its conversion into a first-class operational base. The port and the two existing airfields were to be improved, and three new airfields were to be built in the general Port Moresby area. Figure 4.5.24 Port Moresby in 1942 (McAulay 1991) Coastal Batteries. During the outbreak of war in the Pacific in WWII, several coastal batteries were established at and around Port Moresby, including at Paga Point, Gemo Island, Idler’s Bay and Boera (Douglas 2007:57). Until 1944 several heavy Australian artillery ‘letter’ batteries (named after their designation by letters of the alphabet) were located around the coast at Buna, Milne Bay, Oro Bay and Gili Gili in attempts to repel the advancing Japanese invasion. However, after the fall of several of these areas, the focus of defence shifted to the only remaining base still in allied hands after the Japanese captured New Britain and the north coast of PNG. Port Moresby was a key target for any future enemy operations, as it was an important shipping port for incoming supplies and was needed B. David, B. Duncan, M. Leavesley 4-539 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project to complete the Japanese defence installation network to afford protection to their facilities around the coast (Kidd and Neil 1998:51). Several battery emplacements were built around Port Moresby to augment the Paga battery (built in 1939), including a new battery at Boera. Aircraft Wrecks. Many aircraft crashed around PNG during WWII, either due to being shot down in aerial engagements, through accidental collisions with strafing targets, becoming lost in adverse weather, or running out of fuel. Some aircraft crash sites were relocated, and where deaths had occurred bodies were collected. However, many planes which crashed into the sea or were reported as missing in action were never relocated. Many of these wreck sites became graves for their crews. Several military aircraft are known to have crashed in the Caution Bay area between Boera in the south, Redscar Head in the north and Hidiha Island to the west. These wrecks have been documented by several interest groups, including the Pacific Wreck Database, John Douglas Environmental Consultants and the Department of Modern History of the PNG National Museum and Art Gallery. The Pacific Wreck Database web site was established by a group of avocational (amateur) historians and enthusiasts who have an interest in World War II plane and ship wrecks, and other military installations. This group has documented the background histories of many defence installations and historically known aircraft wrecks around PNG (www.Pacificwrecks.com). Douglas Environmental Consultants have undertaken perhaps the most comprehensive historical studies of the area as part of the environmental analyses for the Konebada Petroleum Park Authority (Douglas 2007). During this time, John Douglas has photographed and documented some military sites in the area (Boera battery), but did not undertake detailed archaeological surveys (see also Pretty 1967). Given the known high concentration of plane wrecks in this region, historical records were investigated to ascertain whether any of these wrecks had occurred inside the Study Area. The identified plane wrecks are listed in Table 4.5.6. Boera Battery. This battery consisted of two 155 mm M1917 heavy artillery guns with associated th range finding station, magazines and other supporting facilities, and was constructed by the 8 Army th Troops Company, Royal Australian Engineers (RAE) (AWM #72247). Construction began on 20 May th 1943 and was completed by 6 July that year, after its guns were transferred from the ‘A’ Australian heavy artillery battery (at Gili Gili, west of Alotau near Milne Bay). The field guns were mounted on a carriage placed over Panama mounts (i.e. a central concrete turret where the tyres of the carriage sat on the turret), and were moved around via the carriage trolley (Plate 4.5.20). The battery was spread over several nearby hills (the southern hill also had a battery Observation Post), was equipped with a range finding station, searchlight, a water well and other support facilities, and was linked to the Coast Artillery Headquarters and Fire Command by telephone and backup High Frequency radio. The battery was sited to cover an unmarked channel into Caution Bay that led through to Port Moresby, which was previously regularly used by local traders (Kidd and Neil 1998:52, 127-29, 339, 341; Douglas, n.d.). Live firing exercises were first undertaken in July 1943. The guns had a firing range of 17 miles, and on one occasion it was recorded that grass fires had been started to the north of the battery by infantry mortar fire, indicating that other military units were also practicing in this area. A Bofors A.A. antith aircraft gun was also installed on 18 July to defend against aerial attack, and was located at the rear of the battery. The battery could be operational in 15 minutes, and the searchlights (with a range of at least 8000 yards) were also intended for use to detect the presence of submarines trying to enter Port Moresby harbour via Caution Bay. Regular practice firing of the guns was undertaken to ensure the battery was ready and competent for any engagement with the Japanese. This included firing at drifting and towed ‘Hong Kong’ targets between 4000 and 14,000 yards away, expending at least 20 (and as many as 40) rounds during each practice session, dependent on the number of officers B. David, B. Duncan, M. Leavesley 4-540 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project present. Rounds were capable of reaching Hidiha/Bavo Islands, and ‘Man-o-war Island’ (to the east of Daugo Island). Anti-aircraft guns also used target balloons for practice. Although an air raid occurred th on 27 September and Japanese planes were heard overhead, no guns were fired by the battery during the incident (Kidd and Neil 1998:129-33). By March 1944, orders were given to prepare to demount ‘A’ unit out of the Boera battery, which was subsequently taken over by ‘O’ unit personnel. When the Boera battery was decommissioned, the guns were returned to Port Moresby and several outbuildings were demolished by Australian contingents who also dismantled the guns from the battery. By July 1944, most of the letter batteries around Port Moresby were removed, as the allied supremacy at that time ensured that the remaining Paga battery was sufficient to repel any attack (Douglas n.d.; Kidd and Neil 1998:52, 134, 340). From 1942-1944, the Hydrographic Branch of the Royal Australian Navy (R.A.N.) began an extensive campaign to map inshore sea routes around New Guinea in a bid to ease the threat to Allied shipping from Japanese submarines and warships in deeper offshore waters (Betty 1996:154-57). In September 1943, the R.A.N. commenced a hydrographic survey of the sea area covered by the battery (Kidd and Neil 1998:131). A chart of unknown date for this area, but probably based on the 1943 survey (Transport-PNG n.d., updated around 1969), shows extensive hydrographic information for the area which is more detailed than modern charts. Plate 4.5.20 #1 Gun, Boera battery in 1944 (photograph courtesy of AWM #072247) Minefields. Captain Charles Kabilu (personal communication 2008), a Sea Pilot with PNG Ports, recalled a submarine net once extending across the reef entrance near Basilisk Beacon on Nateara Reef. Modern charts show that the area to the south of Caution Bay (south of Haidana Island) to Padana Nahua Passage were once mined as part of the former Port Moresby submarine defences (Aus Chart 621). Although these areas have been swept for mines and do not present hazards to surface navigation, the chart notes that the areas are not considered safe for trawling, anchoring or bottoming activities associated with submarines, suggesting that unexploded mines and/or ordinances may still exist in these areas. It is therefore also possible that rogue mines whose positions were not recorded when laid, or which may have become loose and floated away from the minefield and later sunk, may be present in the southern regions of Caution Bay. th Telegraph Cable. On 10 January 1943, a telegraph cable was laid from Cape York (Australia) to Boera Head for the Signals Section (New Guinea Force) and the Post Master General’s Department to link to Port Moresby (AWM photograph #057334). This cable was probably laid as part of the wartime communications network along the PNG coast. The end of the cable was towed ashore via a barge B. David, B. Duncan, M. Leavesley 4-541 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project boat from the steamship Mernoo. From pictures available at the Australian War Memorial web site (AWM photographs #05733-057350 – Plates 4.5.21 and 4.5.22), it appears that the cable was placed ashore in the vicinity of Apau Beach, just north of Boera Point. The cable seems to have been laid directly to the west of the Boera battery. B. David, B. Duncan, M. Leavesley 4-542 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.21 Boera Point, New Guinea. 1943-10-01. Drums with the marine cable attached being towed from the SS Mernoo to the shore (photograph courtesy of AWM #057336) Plate 4.5.22 Boera Point, New Guinea. 1943-10-01. Barge Towing Cable from the SS Mernoo to the Shore (photograph courtesy of AWM #057350) The cable appears to have come ashore just south of the current Telekom radio station which services local communications purposes (Douglas 2007:54), and may explain the later installation of this facility in this area. Although the exact path of the cable across the seabed is not evident from the B. David, B. Duncan, M. Leavesley 4-543 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project photographs, sections of it may lie within the southern portion of the Study Area where the MOF facility is to be situated. The cable was not shown on any charts of the area (Aus Chart 379; Transport PNG around 1969), and no further information is known about this facility at this current time, including whether the cable is still operational. Kidd and Neil (1998:322) have documented that the Boera battery was visited on 2 May 1944 by a Major Curtis to reconnoitre for a submarine cable, which could mean that two cables were installed in the area. New Guinea Forces maps from March 1944 (Galley Reach 3967) also shows that a telegraph cable had been installed from Totoo Village to Redscar Head. This cable was located along the foreshore edge and was submerged at several locations, especially where it crossed streams or rivers. It then proceeded overland from Kido and along the beach to Lea Lea, crossing Lea Lea Inlet and swampland to Papa, before proceeding inland across the Portion 152 area (Lea Lea Inlet 3968 – Figure 4.5.25). Unfortunately, the adjoining map in this series was unavailable, so it is unclear where the telegraph cable linked to at that time. The cable is not shown in later maps of the area (Sheet 5529-IV 1963; Sheet 5130-II, 1965), suggesting that the telegraph cable had either been removed or was no longer used. Although the UNESCO maritime convention (UNESCO, 2001) does not list submarine cables to be of historic significance, the cable’s location would be of archaeological significance due to its associations with PNG/Australian WWII defences. The location of where the cable came ashore would be of historical significance due to its associations with the first telegraph cable between PNG and Australia. Figure 4.5.25 Map 3968: Lea Lea Inlet showing the location of the telegraph cable (shown as dotted line) from Kido to Papa 4.5.9.5 Summary: the Study Area’s European Contact Period Cultural Sites as Identified from Previous Archaeological Investigations After analysing the available historical documentary evidence, oral histories and museum archives and registers, eight types of sites were identified as occurring or likely to occur within or near the Study Area. Many of these occur or are likely to occur underwater: wharves/jetties, underwater telegraph cables, shipwrecks, anchors and anchorages, and some plane wrecks and UXO in particular; a list of the European contact period cultural heritage sites identified from the literature review is presented in section 4.5.10. The site types which we have identified as present, or likely to be present nearby, are: • Shipwrecks, including exploration period ships, coastal traders, and defence vessels. B. David, B. Duncan, M. Leavesley 4-544 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project • Plane wrecks (some of these may be war graves). • Former defence facilities (e.g. landing pontoons/wharves and/or associated pipe bridges over swamp or mangrove areas; battery sites; airstrips). • Possible wharves/jetties associated with former LMS or Cattle Station sites. • Underwater telegraph cables. • Unexploded ordinances (e.g. artillery shells, submarine mines, bombs). • Lighthouses and beacons. • Anchors and anchorages. B. David, B. Duncan, M. Leavesley 4-545 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.10 Survey Results: Site Details This Section presents information on the location of cultural heritage sites within and near the Study Area. Two research methods were followed to obtain this information: • Literature survey. Sites previously recorded by researchers and on record in the register of the PNG National Museum and Art Gallery. This relates mainly to sites near the Study Area, rather than within it, as no previous study of the Study Area has been undertaken. • Field surveys by which cultural heritage sites were directly recorded. The kinds of sites reported include indigenous places (mainly of the pre-colonial period), WWII locations (e.g., plane wrecks, defence installations), early missionary sites and other sites of the early colonial period. For each of the literature surveys and field surveys, these sites are presented separately in two Sections: those on land (Section 4.5.10.1) and those in the shallow marine environment (Section 4.5.10.2). The cultural heritage sites found during the course of the literature reviews (Land: Section 4.5.10.1.1; Shallow Marine: Section 4.5.10.2.1) and during the fieldwork within the Study Area (Land: Section 4.5.10.1.2; Shallow Marine: Section 4.5.10.2.2) are discussed separately. The significance of each site and site complex is addressed in Section 4.5.12, and mitigation measures for managing them are presented in Section 4.5.14 following impact assessments in Section 4.5.13. No pre-European contact period cultural heritage sites have previously been registered by the PNG National Museum and Art Gallery within the Study Area proper. However, based on the literature reviewed in Section 4.5.10 above, a number of nearby sites – some of very high levels of significance (see Section 4.5.12) – have been identified by the Museum and others in the immediate vicinity of the proposed LNG Facilities site area. Ethnographic accounts identify numerous important cultural sites nearby, such as Dirora, Konekaru, Davage and other ancestral villages and legendary places associated with the first lagatoi and the commencement of hiri trade. Within this area can also be found numerous sites dating to the 20th century (in particular WWII sites), along with previously unrecorded marine sites. These previously documented sites have been discussed throughout Section 4.5.7 above. They are listed in Table 4.5.6 (including their Grid References, where known). The archaeological field surveys in the land and shallow marine environments associated with the proposed developments employed in this study have revealed archaeological evidence consistent with the above patterns as revealed from the literature and PNG National Museum and Art Gallery site register. Both pre-European contact and European contact period cultural heritage sites (such as WWII sites) – have been shown to be abundant in the general vicinity of Portion 152. Before presenting the results of the surveys and the individual cultural heritage site details, it is noted that three UXOs were located during the land component of the field surveys, at WGS84 Grid Reference 500991 E 8968212 N (artillery projectile, site CB2a), a UXO at 501934 E 8965139 N (CB2b); and another at 507446 E 8966645 N (CB2c) (Plate 4.5.23). Sub-Lieutenant Mitmit (personal communication 2008) has identified the first of these as a 75 mm (Japanese) artillery shell, which is very similar in size and shape to the other 155 mm projectile heads discovered by Douglas in 2003. Although Mitmit has identified the shell as a 75 mm Japanese artillery shell, it is more likely that it is a 155 mm fired from the Boera battery. The fuse mechanism works by delaying the explosion of the shell until a preset time determined by the dial up mechanism on the head of the shell. Sometimes the shells used a chemical or battery inside the fuse. This shell had an electrical fuse, with the propellent charges still located inside the main body for the shell. Petty Officer Yamun (personal communication 2008) stated that as one shell had been found in the area, it was likely that there were many more. The projectile was located about 5.5 km from the Boera battery, and is likely to have been fired from that location. B. David, B. Duncan, M. Leavesley 4-546 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.23 Left: Lt Tui Gaileko, Petty Officer Steven Yamun and Leading Seaman Ausa Tau with unexploded artillery projectile discovered in the LNG152 area during the terrestrial archaeological survey. Right: Artillery projectile head discovered in LNG152 area. Note armour piercing head and fuse are intact, indicating the projectile was live Plate 4.5.24 WWII artillery projectiles held at Boera village (from Taylan, 2003) Discussions were held with the PNG Bomb Squad to determine the extent of WWII unexploded ordnance (UXO) material previously discovered in the Study Area. Lt Tui Gaileko and Petty Officer Steven Yamun (personal communication 2008) reported that the squad had previously been called to B. David, B. Duncan, M. Leavesley 4-547 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project remove or disarm unexploded ordnances in the southern parts of the LNG152 area, and in particular large 500 lb bombs had been detonated by them close to Boera, an observation also reported by Boera villagers (Mea Dobi, personal communication 2008). Unexploded ordnances were previously discovered on-shore between Papa and Boera by women searching for mud crabs who had seen artillery projectiles in the swamps (CB2b). Bombs and artillery projectiles (CB2c) had also been found on the eastern side of the Papa road just below the ranges (Gau Ario and Daro Avei, personal communication 2008). No ordnance was reported in the waters adjacent to these areas by the bomb squad, although local villagers (Gau Ario, personal communication 2008) maintained that they were aware of ‘shells’ which had been located underwater and subsequently blown up by the bomb squad. The presence of the sea minefield to the south of Caution Bay (see Section 4.5.9.4 for details) also raises the possibility that rogue mines, whose positions were not recorded when laid or which may have become loose and floated away from the minefield and later sunk, may be located in the southern regions of Caution Bay, although none were found during our surveys. UXOs are recorded here as they are technically archaeological sites relating to the region’s history, as well as to report them for safety reasons. Several 155 mm artillery projectile heads with intact detonators were discovered in Boera village in 2003, which had been collected by villagers possibly for their scrap metal value (Douglas, no date) (see Plate 4.5.24). These UXO issues are mentioned from the onset, but are not discussed further in this report. 4.5.10.1 Onshore LNG Facilities Component A review of previous archaeological research on indigenous sites undertaken on land in the broader Port Moresby region indicates ‘a small number of extensive sites, and a larger number of small sites’ along the coastal zone, as well as sites on the inland plains and on the ‘summits and upper slopes of the range of inland hills lying between the [Laloki] river and the coast (Bulmer, 1978:71) (see Section 4.5.7). Many of these elevated inland sites, as well as some offshore sites, were located in such landscape settings for defensive purposes (see also Orrell, 1977:15). Significantly, Bulmer (1978:72, emphases added) concludes that: There seems to be evidence for nucleated settlement in the presence of very large prehistoric sites at a few places along the coast. Four have been located so far [in the Port Moresby region], but it is possible that others will be found. This can be expected because there appears to have been discontinuity of settlement at the large sites known, so the large settlements may have been periodically relocated. … It also seems that settlement on the inland river plains was nucleated, with concentration of habitation sites at only two centres and very little indication of occupation of any scale elsewhere. This may be obscured by the erosion and sedimentation following forest clearance. It may also be expected that other centres of occupation on the river plains may be found. Bulmer (1978:72-73) also notes that ‘There appear to have been changes in the position of the coastline during the period of human occupation’ caused by increasing alluvial and colluvial siltation resulting from forest clearance with the onset of hinterland gardening. Changing coastlines in the past have resulted in a seaward progradation of the coast. Earlier sites can thus be expected to be buried beneath sediments in what is now the hinterland (but in the past would have been ancient shorelines). Bulmer (1978:73) thus notes that ‘some of the sites now some distance inland were much closer to the coast in the past. Thus a certain amount of settlement described … as “inland” may in the past have been coastal’. These environmental changes ‘affected the position of coastal settlement. The most substantial changes have probably occurred in the far western part of the survey area’ – that is, including the Lea Lea-Boera area incorporating the Priority Area of the present report. B. David, B. Duncan, M. Leavesley 4-548 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.10.1.1 Literature Survey Indigenous Cultural Heritage Sites Bulmer’s systematic study of a small sample of archaeological sites in the broader Port Moresby region indicates an absence of large ceramic sites inland, but their presence in the coastal and nearcoastal zone. She further notes that ‘Not a single pre- or non-ceramic site has yet been found. The earliest ceramic site so far dated is only about 2,000 years old, so it seems obvious that we have more yet to find’ (Bulmer, 1978:62-64). That is, despite her study the earlier phases of human occupation in the Port Moresby region have not yet been found, although such sites are expected to occur given that we know that people have been in PNG for tens of thousands of years. Bulmer (1978:63) further notes that ‘Allen and Swadling have both suggested … that the south coast of Papua was essentially unoccupied prior to the earliest ceramic horizon that has so far been found’ – i.e., around 2000 years ago – but we now know this was not the case due to the recent discovery of terminal Pleistocene (pre10,000 year old) and early Holocene archaeological materials at sites OJP and Wokoi Amoho (respectively) in the Gulf Province lowlands (David et al., 2008). Bulmer (1978:64) states that 47 of the 95 settlement sites recorded in the broader Port Moresby region by 1972 were located on or near the coast, while the other 48 were inland (‘i.e., not on the beach or on the hills above the beach or in small coastal valleys’). Apart for four very large sites interpreted as nucleated village sites, all other reported coastal archaeological sites were small, indicating either cultural places of limited spatial extent or once-larger places that have suffered considerable erosion. These smaller coastal sites typically contain pottery sherds, stone artifacts, oven stones and/or shells (Bulmer, 1978:66). In total 11 coastal settlement sites had been recorded to the west of Port Moresby itself at the time of Bulmer’s writing: AAV at Cogland Head; AAN, AAS, AHQ, AHR and AGI at Idlers Bay; AFD and AFR at Ranvetutu; AAO between Ranvetutu and Koderika; AAR at Porebada; AAP and AAQ at Daugo Island; and ABH, ABI, AEQ, AFA, AFB, AMG, AMH near Boera (Bulmer, 1978:68; see also Lampert, 1968). ‘All of these sites are thin scatters of midden, except the AFD-AFR sites at Ranvetutu, and AFA and AMG-AMH at Boera. These are very large deposits covering acres of ground’ (Bulmer, 1978:68). Four smaller coastal sites on estuarine plains (ABZ, AHV, AHW and AHZ) were recorded near Lea Lea village ‘on the only two islands of elevated land on this stretch of the coast’ (Bulmer, 1978:69). She concludes that ‘at some point in prehistoric times nearly every suitably protected coastal site has been occupied, but only a few were repeatedly or continuously inhabited’ (Bulmer, 1978:64-65). That is, in many cases coastal settlements were relatively short-lived as some villages and other site types were abandoned while others became established. One such major archaeological site is AMH (Ava Garau, see Section 4.5.8.4 and Section 4.5.8.7) about 1 km westnorthwest of the present village of Boera, which Bulmer (1978:65), following Swadling (n.d.), notes is ‘much larger’ than 30,000 m2 and ‘perhaps 1,000 m long, with other very large sites also nearby’. This represents one of the largest – and probably the largest – known archaeological site in the entire greater Port Moresby coastal region, with only Taurama (30,000 m2), Motupore (20,000 m2) and Ranvetutu (site AFD, about 3 km south of Porebada: 9,000 m2) of comparably impressive size (Bulmer, 1978:65). Of the 48 settlement sites identified by Bulmer from the hinterland, 28 come from the coastal hills (AAE-AAJ, AAX, ABB, ABJ, ABN-ABO, ABU, ABY, ADD, AFJ, AFN, AFO, AHG, AIC-AIK and AKU). Bulmer (1978:69) notes that these are ‘nearly all sites that could be interpreted as defensive’. ‘The sites in the coastal hill zone consist almost entirely of midden deposits on the summits or high slopes of the highest range of hills’, although some contain ‘large areas of pottery … and could reflect settlements of substantial size’ (Bulmer, 1978:69). Shortly to the northwest of Boera, and largely addressing the same sites that Bulmer discussed (above), Swadling (1980:104) also identified a ‘huge archaeological complex which includes Davage, the legendary village from which oral traditions relate the first hiri … was made’ and the home of Edai Siabo (see also Swadling, 1977:40; see Chatterton, 1968:93 for discussion of Davage as an important B. David, B. Duncan, M. Leavesley 4-549 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project origin village for the people of Boera) (Figure 4.5.26). This site complex includes Ava Garau (site AMH), an ancient archaeological village site containing evidence of Bulmer’s Style I (Swadling’s Early Period) ceramics and dated to 1220±95 BP (for details, see Section 4.5.8.4 and Section 4.5.7.10 above). Elsewhere she (Swadling, 1977:39, italics added) notes that ‘a huge village was founded at Boera [Ava Garau] about 1,200 years ago. The broken pottery and shells there cover the largest area of such village rubbish reported anywhere between ManuManu and Gabagaba. Many people must have been living there 1,200 years ago. It was probably the most important village in the Port Moresby area’. She writes (Swadling, 1976:8): Ava Garau is the name given to one of the hills forming a sea cliff ridge which runs for more than a kilometre from the west of the present Boera village. This cliff ridge is broken by two seasonal streams which give easy coastal access. Apart from these cliffs the flat floor valley has no defensive features. Milne Bay-like pottery is found along the entire inland side of this cliff ridge and small scatters are present on the large Nemu ridge further inland. The entire complex is the largest area of comparable ceramics reported in the Port Moresby area. All in all suggesting a considerable centralisation of settlement in the Port Moresby region at Boera at this time. Figure 4.5.26 Boera sites having sherds with designs similar to those on late Red Slip sites (courtesy of the PNG National Museum and Art Gallery) Swadling (1977:40) suggests that upon abandonment the people of Ava Garau may have founded Motupore, Taurama and other major sites further to the east, significant ancestral village sites of oral tradition. This further re-affirms the presence in the Lea Lea – Boera area of sites covering at least 1200 years – and possibly considerably more – of Motu/Koita history. As Swadling (1977:41) notes, ‘The Boera people have been shown … to have been in that area for a long time’. B. David, B. Duncan, M. Leavesley 4-550 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project At some distance from Portion 152, but also of great cultural significance to the people of Lea Lea and other local villages, is the ancestral settlement site of Buria. Buria is a hilltop village site located some 6 km north of Lea Lea (this hill is ‘incorrectly labeled “Darebo Hill” on the Redscar Head map – 5130II’, Johnston, 1973:1). While the hill itself is of cultural significance, little archaeological or anthropological documentation of this site has been undertaken. Gabrielle Johnston (1973:1) undertook rapid reconnaissance surveys of the hilltop and slopes, where she confirmed the presence of ‘large deposits of pottery fragments, shell, and stone’ at site AHW. Here she recorded the presence of ‘late’ Motu pottery, as well as more ancient red slip wares of a kind typical of the earliest ceramic phase in the Port Moresby region, suggesting ‘a considerable antiquity for the site’ (Johnston, 1973:2). A traditional dancing ground was recorded some 500 m away, and at site ANN nearby is found the place where a surprise attack had forced the people of Buria to abandon their village (see below). Sere (1975:81, 86) records a traditional story about a massacre at Buria village, identifying the event as having taken place around 300 years ago by genealogical reckoning: There was a relationship between Buria and Pari which resulted in a clan-leader’s son from Buria being betrothed to a clan-leader’s daughter from Pari. However the Buria boy preferred to marry a girl from Daeroto and the girl was therefore married to another Buria boy who was not a clan-leader’s son. The Pari wife therefore lived at Buria feeling insulted, and when she went to Pari to visit her family she complained about how she had been down-graded and added the information that the Buria were boasting that they were the best dancers among the Motu-Koita. Also she described Koita customs in such a way that the Pari people determined to attack Buria. The Pari successfully surrounded and attacked Buria while the latter were dancing. The bows and arrows of the Pari wrought devastation among the Buria and (according to legend) the only escapee was Goroa who hid in one of the Buria caves. It was over the Daeroto land that war was to break out between the Lea Lea and the Doura. Further archaeological sites were found in the Buria area, including site ANO ‘which covered a similar area to that of AHW’ (see Plate 4.5.25) (Johnston, 1973:3). Other nearby archaeological sites include AHV and AHZ. Clearly this entire area is of great cultural significance and would need to be avoided when planning future developments associated with the LNG processing facilities. B. David, B. Duncan, M. Leavesley 4-551 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.25 View of Buria village hill from the sea south near Papa looking north beyond Rea Rea (from Johnston, 1973:4) Additional to the above settlement sites, clay sources are also known from the general vicinity of the Study Area. Groves (1960:8-11) notes that the Manumanu villagers did not possess clay sources near their own village, and therefore sent parties to extract clay for pottery-making from sources ‘at a site near the beach between ReaRea [Lea Lea] and Papa’. He further notes that while for Manumanu villagers ‘individuals may dig clay at any time of year’ and ‘no special rules govern access to the clay’, this is not the case for clay obtained specifically for hiri pots whereby ‘custom prescribes that only members of the [lagatoi’s] … crew may fetch clay’; generally the women dig the clay from pits around 1 m deep, except during these special hiri occasions when it is only men who obtain the clay and bring it to the women for the manufacture of pots (Groves, 1960:11). However at Boera where other sources of clay occur, ‘each woman fetches clay as she pleases’, irrespective of whether she intends to make pots for domestic use, regional exchange or hiri trade, and here no ritual observations are followed (Groves, 1960:14). Worthing (1980) has identified clay sources near Boera, and at Boki between Lea Lea and Papa, in line with Groves’ own observations. The clays and sand tempers that were used in archaeological ceramic sherds were identified and matched against modern source samples and thereby used to track back in time the history of hiri trade ceramics. Boera clays were used to manufacture pottery that has been found as far west as the Gulf Province, eastward to Motupore Island (Worthing, 1980:87), while Boki clays were used in pottery found from the Gulf Province to the Papa Salt Pan site (AWL). Worthing (1980:92-93, 98) notes that in 1980 the Boki clay pits were still used by Lea Lea potters ‘who mix it with Lea Lea beach sand’. This Boki clay source has been recorded as PNG National Museum and Art Gallery site AFC. The Boera clay source has been recorded as site ABG (see Table 4.5.6 for summary details). While traditional freshwater wells are known to occur near Boera, these have not yet been recorded. The PNG LNG Project developers should nevertheless be alert as to their presence, as their exact locations and significance cannot be addressed in this report due to their occurrence some distance outside the Study Area (see Table 4.5.6 for summary details). At the time of writing, and incorporating the sites listed above, the PNG National Museum and Art Gallery register along with various historical documents as discussed in Section 7.0, contained the following archaeological sites for the region shortly surrounding the Study Area (apart for the Buria sites which lie further to the north, this list is bounded by the Port Moresby 1:50,000 topographic map Grid References (in AGD66) (for GPS points see original study – David 2008) (thus covering a land area >3 km on all sides of the Study Area; see Table 4.5.6 for locational and other details of each site): • AADI, AAGM, ABG, ABH, ABI, AEZ, AFA, AFB, AHW, AHY, AMG, AMH, AMI, ANA, ANN, ANO, ANU, AOG, AOH, AOI, AOJ, AOK, AOL, AOM, AOX, APC, APF, APG, ARD, ARE, ARF, ARG, ARH, ARI, ARJ, ARK, ARL, ARM, ARQ, ASM, AWL (see Figure 4.5.27). We note that Graeme Pretty found three sites on Stanley Hill (Uda Bada on the Port Moresby 1:50,000 topographic map) in 1967; these sites appear to be represented by those listed in the PNG National Museum and Art Gallery list presented here (as indicated by the Museum’s register map rather than its site survey sheets), but it is not clear from these records which of the museum codes represent the Pretty sites. In addition, the following oral tradition sites mentioned in the historical records can be added, as outlined in Section 4.5.7 (see Table 4.5.6 for further details): • Buria (includes ANN), Davage, Aemakara, Daeroto, Konekaru, Dirora, Taubarau, the site of Edai st Siabo’s 1 lagatoi anchor (ASM), the Boki clay source (AFC), the Boera clay source (ABG), the Boera freshwater wells. B. David, B. Duncan, M. Leavesley 4-552 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Of these oral tradition places, only AFC, ABG, ANN and ASM have been located and given formal site codes by the PNG National Museum and Art Gallery, although archaeological sites at Davage (ANU), Buria (AHW, ANN, ANO) and Taubarua (AMG) have also been recorded. Figure 4.5.27 Location of archaeological sites between Papa and Boera on the PNG National Museum and Art Gallery site register and register map The broader Lea Lea–Porebada area surrounding the Study Area undoubtedly contains a very large number of archaeological and oral tradition sites that remain undocumented, and therefore cannot be presented in this report, because they lie outside the cultural heritage study brief (thus, outside the B. David, B. Duncan, M. Leavesley 4-553 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Study Area that the cultural heritage survey team was asked to cover in this report). Some of these sites are likely to be very important. While the sites listed here (see Table 4.5.6) thus document the cultural heritage places uncovered during the literature review, they are only an opportunistic coverage of the sites that lie outside the Study Area, and are therefore not meant to represent an exhaustive or systematic review of sites that exist outside this area. Systematic field surveys would be needed for such a task. While some of the archaeological sites listed above correspond geographically with some of the oral tradition sites, the two are not synonymous nor entirely overlapping, as individual archaeological sites tend to be relatively small locations delimited by the spread of anthropogenic material objects on the landscape, whereas the oral tradition sites tend to not be delimited simply by the spread of material objects made or dropped on the ground by people in the past. A case in point is the oral tradition site of Buria, which essentially consists of the hill of that name and its immediate surroundings rather than the isolated archaeological sites AHW, ANN and ANO which occur there (despite some limited opportunistic mapping of archaeological sites such as AHW, ANN and ANO, the Buria area has not yet been systematically mapped for oral tradition or archaeological sites, although the oral tradition site complex is summarily recorded below as site CB10 – see Section 4.5.10.1.2). Locational and summary details are presented for each archaeological and oral tradition site in Table 4.5.6. These sites are not discussed further here for they occur outside the Study Area, although they will be further considered in Section 4.5.12, Section 4.5.13 and Section 4.5.14 when issues of significance, impacts and management recommendations are made. WWII Plane Wrecks Two WWII plane wreck sites occur near the Study Area. One of these is a P-39 (P-400) Aerocobra been recorded by members of the cultural heritage site surveying team in the field as site CB4. The other is a P-38 Lockheed Lightning which crashed on the Lea Lea airstrip in 1944. A third plane, an A-6M2 Model 21 Zero, crashed near Boera in 1942 (it is not known whether this plane crashed on land or water). The whereabouts of this site is unknown. Details of these three sites are presented in Table 4.5.6. 4.5.10.1.2 Field Surveys The above details relate to cultural heritage sites revealed through a search of the literature. The present Section now presents the results of the field surveys within the Study Area, and of limited opportunistic visitation to sites outside but near the Study Area. B. David, B. Duncan, M. Leavesley 4-554 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Oral Tradition Indigenous Cultural Heritage Sites AADI Boera Midden AAGM A beach site excavated by Les Groube in 1988 and analysed by Gillian Cox. Contains pottery sherds, flaked stone artifacts, animal bones, shells and worked shell bracelets. Excavation results have never been published. x This site is where a stone club was found in a garden shortly to the east of Papa village; the Grid References appear to have been incorrectly entered as ‘706705’ on the PNG National Museum and Art Gallery register (but appear to be correctly entered at 5015 E 89705 N on the map). Correct GPS points available in original report – see David (2008) (field checking required). A stone club was found in the gardens at this site. x ABG Boera Clay Source Grid References estimated from PNG National Museum and Art Gallery register map only; site register form not seen. Bulmer (1978:44) identifies this site as a clay source near Boera village (see Section 4.5.8.1.1). x x ABH Raroasi This beach village site appears to have been incorrectly mapped as ‘001635’ N on the PNG National Museum and Art Gallery register; on the register map it is at 5018 E 89636 N; it needs to be correctly remapped. Correct GPS points available in original report – see David (2008). The PNG National Museum and Art Gallery site survey record for this site states ‘Pre-war village around point to north of present village. After war-time disposal, village re-assembled at present site’. x x ABI This large midden is a beach site. It has been incorrectly mapped as ‘033611’ N on the PNG National Museum and Art Gallery register; on the register map it is at 5038 E 89611 N; it needs to be correctly remapped. Correct GPS points available in original report – see David (2008). This is a large midden that includes ceramic sherds and stone artifacts. B. David, B. Duncan, M. Leavesley 4-555 x Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Indigenous Cultural Heritage Sites (cont’d) Grid References estimated from PNG National Museum and Art Gallery register map only; site register form not seen. AEZ AFA Grid References estimated from PNG National Museum and Art Gallery register map only; site register form not seen. x x AFB This hilltop site behind Boera Head has been incorrectly mapped as ‘007618’ on the PNG National Museum and Art Gallery register/map; it needs to be correctly re-mapped. Correct GPS points available in original report – see David (2008). The site contains pottery sherds and stone artifacts. AFC Boki? PNG National Museum and Art Gallery site register and map not seen; the Grid References are therefore not known, although site consists of two places located 100 m apart on the beach between Lea Lea and Papa. This is almost certainly the Boki clay source (see Section 7.1.1). AHW Buria Located to the north of Lea Lea village; site register details not accessed. Site contains a scatter of pottery sherds and shell over an area ~60’ x 100’. x AHY This site is located in a creek bed; the Grid References appear to have been incorrectly entered as ‘035639’ on the PNG National Museum and Art Gallery register, but appear at 5040 E 89639 N on the map. Correct GPS points available in original report – see David (2008) (field checking required). A stone club-head and a stone axe blade were collected from this site in 1972 and 1996 respectively. x AMG This site is located on the inland side of Iduata or Taubarua hill; the Grid References appear to have been incorrectly entered as ‘007623’ on the PNG National Museum and Art Gallery register. Correct GPS points available in original report – see David (2008) (field checking required). This is an extensive deposit of midden pottery, shell and stone artifacts. X Taubarua B. David, B. Duncan, M. Leavesley 4-556 x Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Oral Tradition Indigenous Cultural Heritage Sites (cont’d) AMH Ava Garau This site is located on the inland side of Ubo and Ava Garau hills; the Grid References appear to have been incorrectly entered as ‘004628’ on the PNG National Museum and Art Gallery register, but appear to be correct on the map. Correct GPS points available in original report – see David (2008) (field checking required). This major archaeological site contains an extensive deposit of midden pottery, shell and bone. This site was excavated by Pamela Swadling, who obtained a radiocarbon determination. X Lauara Koupana This site is located on the inland slopes of Lauara Koupana; the Grid References appear to have been incorrectly entered as ‘001634’ on the PNG National Museum and Art Gallery register (but appear to be correct on the map). Correct GPS points available in original report – see David (2008) (field checking required). The site extends from the valley floor rd up the slopes of the hill (to about 1/3 up the slope). It contains a great concentration of pottery sherds. X ANA Nadibada Grid References estimated from PNG National Museum and Art Gallery mud-map, and corroborated with their site register map; estimate only. The PNG National Museum and Art Gallery site survey record was not accessed. X ANN Buria Located on the summit ridge of Buria to the north of Lea Lea village; site register details not accessed. It is said to be a sing-sing area for the old Buria village. ? ANO Buria Located to the north of Lea Lea village; site register details not accessed. It contains a sparse scattering of plain sherds (including a near-complete cooking pot). A freshwater hole is close by. X AMI B. David, B. Duncan, M. Leavesley 4-557 x Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) Comments* Type of Cultural Heritage Site Archaeological Oral Tradition Grid References estimated from PNG National Museum and Art Gallery mud-map, and corroborated with their site register map. The PNG National Museum and Art Gallery site survey record was not accessed. x x AOG This site includes pottery sherds. x AOH Site recorder is not certain of exact Grid References; estimate only. This site is on a flat saddle between hill with prominent stone block formation and Nemu. It includes pottery sherds. x AOI Site recorder is not certain of exact Grid References; estimate only. This site is on flat slight saddle area between the hill with prominent stone black formation and Nemu. It contains pottery sherds. x E N Indigenous Cultural Heritage Sites (cont’d) ANU Davage AOJ Site recorder is not certain of exact Grid References; estimate only. This site is on the lower slopes of the Nemu war-time fortified summit. It contains pottery sherds. x AOK Site recorder is not certain of exact Grid References; estimate only. This site is on the mid-slopes of the Nemu war-time fortified summit. It probably contains pottery sherds (uncertain from PNG National Museum and Art Gallery site survey record form). x AOL Site recorder is not certain of exact Grid References; estimate only. This site is on the summit of Nemu, and has been greatly disturbed by the construction of gun implacements. It’s contents include pottery sherds. x AOM The PNG National Museum and Art Gallery site survey record for this site does not specify site contents. x AOX This site is on a hill. It contains unspecified archaeological materials. x The PNG National Museum and Art Gallery site survey record for this site does not specify site contents but states ‘old settlement site Boera area’. x APC Vaiboda B. David, B. Duncan, M. Leavesley 4-558 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Oral Tradition Indigenous Cultural Heritage Sites (cont’d) APF Iva This site has an army trench on ridge up another summit. Its archaeological contents are not apparent from the PNG National Museum and Art Gallery site survey record. x APG Ugava This site is considered by Koderika people to be a garden camping place. It includes shell midden material and glass. x Site recorder is not certain of exact Grid References; estimate only. Based on pattern of errors on PNG National Museum and Art Gallery register/map, and sketch map on site recording form, the correct Grid References are probably approximately those presented here. This site is a scatter of pottery sherds. x ARD ARE The site consists of two small scatters of pottery sherds on the eastern side of the back mudflats of the Vaihua River inlet. Based on pattern of errors on PNG National Museum and Art Gallery register/map, and sketch map on site recording form, the correct Grid References are probably approximately those presented here. x ARF The site consists of two small scatters of pottery sherds on an old sandspit of the Vaihua River inlet. x ARG This site is a scatter of pottery located on an old sandspit along the Vaihua River inlet; the Grid References appear to have been incorrectly entered as ‘019656’ on the PNG National Museum and Art Gallery register. Correct GPS points available in original report – see David (2008) (field checking required). The site is a scatter of pottery sherds. x ARH This site is a scatter of pottery located on an old sandspit; the Grid References appear to have been incorrectly entered as ‘022657’ on the PNG National Museum and Art Gallery register. Correct GPS points available in original report – see David (2008) (field checking required). The site is a scatter of pottery sherds. x B. David, B. Duncan, M. Leavesley 4-559 x Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Indigenous Cultural Heritage Sites (cont’d) ARI This site is a scatter of pottery located on a small remnant island; the Grid References appear to have been incorrectly entered as ‘017663’ on the PNG National Museum and Art Gallery register. Correct GPS points available in original report – see David (2008) (field checking required). The site is a scatter of pottery sherds. x ARJ This site is a small scatter of pottery sherds on lower slopes above mudflats of the Vaihua River inlet. Based on pattern of errors on PNG National Museum and Art Gallery register/map, and sketch map on site recording form, the correct Grid References are probably approximately those presented here. x ARK This is a small scatter of pottery sherds and shells on a large sandspit at eastern mouth of Vaihua River inlet. Based on pattern of errors on PNG National Museum and Art Gallery register/map, and sketch map on site recording form, the correct Grid References are probably approximately those presented here. x ARL A small scatter of pottery and shells. Based on pattern of errors on PNG National Museum and Art Gallery register/map, and sketch map on site recording form, the correct Grid References are probably approximately those presented here. x ARM This site is on a sandspit in from the mangroves, along the western mouth of Vaihua River inlet. It contains three small scatters of pottery. Based on pattern of errors on PNG National Museum and Art Gallery register/map, and sketch map on site recording form, the correct Grid References are probably approximately those presented here. x B. David, B. Duncan, M. Leavesley 4-560 Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Oral Tradition Indigenous Cultural Heritage Sites (cont’d) ARQ st This site is located in a creek bank; the Grid References appear to have been incorrectly entered as ‘048628’ on the PNG National Museum and Art Gallery register, but appear at 5055 E 89629 N on the map. Correct GPS points available in original report – see David (2008) (field checking required). A few sherds were eroding out of the creek bank in 1979. A radiocarbon sample was undertaken for this site by Colin Pain. The result does not appear to have been published. x According to the PNG National Museum and Art Gallery site survey record, this site contains a fragment of granite boulder said to be a portion of the anchor-stone of the first boat to reach what is now Boera. The site also contains a round stone, stone tools and stone pottery. x This site is in the middle of a flood-prone flat area opposite the road from Papa village. It contains ceramic sherds. An archaeological excavation was undertaken here by Pamela Swadling, and a radiocarbon date has been published. x ASM Edai Siabo’s 1 lagatoi anchor AWL Papa Salt Pan CB10 Buria This is the ancestral village hill-site of oral tradition. It remains to be systematically recorded. x Davage This important cultural heritage site just north of Boera has not been recorded (but see archaeological site ANU). x Aemakara This extremely important cultural heritage site is an ancestral village to the Koita. The site is very extensive in size, estimated to measure approximately 1.0 x 0.7 km in area. The GPS reading given here is for the site’s high point; Aemakara continues westward from the hill-top, along the slopes to the mudflats. A rich archaeological assemblage can be seen here, corresponding with the area of Aemakara as known from oral traditions. At the foot of the hill whose GPS location is given here, archaeological deposits include extensive pottery and chert artifacts. Daeroto This important cultural heritage site has not been recorded. B. David, B. Duncan, M. Leavesley 4-561 x x x x Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Oral Tradition Indigenous Cultural Heritage Sites (cont’d) Konekaru This past village site consists of the area encompassing CB1, JD1, JD2, JD3, JD5 and ML3, including Konekaru beach from its northern to its southern ends. The present entry concerns the cultural place of Konekaru village; it contains varied archaeological expressions which have been recorded separately as a number of distinctive archaeological sites. x Dirora This is a very important ancestral Koita village site (also known as Namura), located approximately 6 km NE of Boera. This important cultural heritage site has not been recorded. x Taubarau This important cultural heritage site just north of Boera has not been recorded (but see archaeological site AMG). x Boera Freshwater Wells This important cultural heritage site near Boera has not been recorded. x European-contact period cultural heritage sites CB30 Boera Battery Sisal Farm/Schimmer Airstrip/Fairfax Cattle Station This is a major defence facility near Boera. It has not been systematically recorded. x During biological surveys associated with Konebadu Petroleum Park Venture, Douglas (2007:54) reported that little was left of the former cattle station beyond a derelict windmill and remnants of a building, and almost nothing was visible of the former airfield. Douglas hypothesised that a gap in the mangroves just outside the former station (at Konekaru beach) was the result of clearance for the construction of a (possible) former wharf on this site used to ship sisal. While an individual archaeological site relating to these European-contact period activities has been recorded during the field surveys (site JD4), the Sisal farm/Schimmer airstrip/Fairfax cattle station themselves have not been recorded as historical sites. x B. David, B. Duncan, M. Leavesley 4-562 x Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Aircraft wrecks and other WWII sites P-38 Lockheed Lightning – Crashed Lea Lea Airstrip (Portion 152 Area) P-38 Lockheed aircraft wreck which crashed at Lea Lea Airstrip. Now housed at the DoMH Museum. A Lockheed P-38 Lightning aircraft (piloted by Captain Wayne Rodherb th th of the 39 Squadron, 35 Fighter Group) crash-landed on the Lea Lea (or Schimmer) airstrip in 1944 during an emergency landing when it ran out of fuel. Due to its low profile and proximity to local swamps, the ground was soft and often waterlogged such that the airfield was considered as an emergency landing strip only. These were the conditions that led to the wreck of the P-38. To enable other aircraft to use the facility, the P-38 was pushed to one side of the strip and left in situ. The aircraft and all its parts were later recovered by PNG Defence Force personnel in 1978. It is now stored at the DoMH Museum’s premises in Boroko (Mark Katakumb, personal communication 2008; DoMH display). P-38 Lockheed Lightning aircraft of the type used in the pacific war (from Morse, 1984:1427). B. David, B. Duncan, M. Leavesley 4-563 x Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Aircraft wrecks and other WWII sites (cont’d) CB4 P-39 (P-400) Airacobra (Serial # AP378) near Boera (Vaihua/Nadi Vasiga) P-39 Airacobra aircraft of the type used in the pacific war (from Morse, 1984:1428) A P-39 (P-400) Airacobra was shot down and either crashed or forced to land on the coast near Boera on th 4 July 1942 whilst intercepting a formation of Japanese aircraft. The fighter-plane was piloted by nd 2 Lt Frank Angier who belonged to the 39th Fighter th Squadron, 35 Fighter Group and who survived the crash. The location was recorded as 30 miles from Port Moresby (http://www.pacificwrecks.com/aircraft/p400/AP378.html; Air Museum, no date:192). John Lelai (personal communication 2008) of DoMH (PNG National Museum and Art Gallery) has inspected this site, although no report was available from DoMH. At the time of inspection by the DoMH, the fuselage and one engine and propeller were relocated in the mud near the Vaihua River. The site was accessible via a track on the southern side of the river. Douglas and Sawanga (2007:54) reported that this aircraft crashed into the mangrove swamps just north of Boera close to the second land road parallel to the coast. In 2007 the Boera people showed Douglas (personal communication 2008) a 0.5 calibre machine gun with a wooden handle which had come from the plane wreck. B. David, B. Duncan, M. Leavesley 4-564 x Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Archaeological Aircraft wrecks and other WWII sites (cont’d) CB4 (cont’d) Type of Cultural Heritage Site During the oral history interviews, Papa and Boera residents reported the remains of a plane in the mangroves between these two villages. Although the plane was regularly seen by women collecting mud crabs, they avoided the site as they believed it was a place where spirits dwelt and hence they were afraid to go there (Daro Avei, Gau Ario and Moi Dobi, personal communication 2008). Boera villagers reported that in recent years large Section s of the plane have been salvaged for its scrap metal value, leaving only the engine and cockpit frame in situ (Moi Dobi, personal communication 2008). This site was not inspected due to landward access restrictions imposed by the discovery of unexploded ordnances in the LNG Facilities site area. There is also some confusion as to the exact location of the site. John Lelai and Dobi Avei (personal communication 2008) place it near Viahua River (location CB4a), whereas Moi Dobi (personal communication 2008) stipulates that it is closer to Nadi Vasiga (location CB4b) near the proposed location for the Marine Offloading Facility. During the surveys, Matthew Leavesley was taken closet to a WWII plane crash site in the mangroves at location CB4c by local villagers; this is likely to be the site of the plane crash discussed here, and thus represents a third alternative location. It is imperative that this site be relocated to ascertain its exact location before any works begin at the Marine Offloading Facility site. B. David, B. Duncan, M. Leavesley 4-565 Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Aircraft wrecks and other WWII sites (cont’d) CB21 This Imperial Japanese Navy aircraft was built by th Mitsubishi on 14 January 1942, and crashed near Boera. Although the crash site was surveyed by the th Allied Transport Investigation Unit (ATIU) on 11 July 1942 (referred to as site AD-10), no other information was located as to the whereabouts of this site (http://www.pacificwrecks.com/aircraft/a6m2/3537.ht ml). It may be archaeological site CB21, estimated to be located at the Grid References listed here (Grid References are approximate only). A-6M2 Model 21 Zero (Manufacturer # 3537) – Boera x A-6M2 Nakojima Seaplane (from Morse, 1984:2434) th B-24 Consolidated Liberator USAAF – Boera B-24 Consolidated Liberator (from Morse 1984:155) On 8 August 1943, a B-24 aircraft was reported to have crashed in the mangroves near Caution Bay after the crew bailed out. It was reported that the aircraft had been circling Boera in heavy overcast weather around 9 pm before crashing. The next day USAAF personnel commenced a search to relocate the plane, which was finally found (with a body inside) by Fortress Signal Corps personnel attached to the Boera battery. The find was reported to USAAF (Kidd and Neil, 1998:130). There are no reported specific site locational details for this site. B. David, B. Duncan, M. Leavesley 4-566 x Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Aircraft wrecks and other WWII sites (cont’d) CB28 B-25D Mitchell Bomber (Serial # 4130496) – Bava Island A B-25 Mitchell Bomber (from Morse, 1984: 829) On 31 July 1943, a B-25 aircraft was reported to have crashed into the sea about 8 miles at 190º from the Boera battery (Kidd and Neil, 1998:130). The st Mitchell bomber (serial) was part of the 71 Bomb th Squadron, 38 Bomb Group, and departed Durrand Airstrip to undertake a practice bombing run on the wreck of the S.S. Puth on Manubada Island (south of Port Moresby). However, during the low level bombing practice in which the bombs were skipped across the water surface, the bomb bounced higher than expected and the resulting explosion damaged the engines and tore off Section s of the wings. The plane remained airborne but uncontrollable for eight minutes, before crashing onto Bava (Bavo) Island killing all on board. A small launch sent to retrieve bodies the next day sank after exploding due to fuel in the bilges, whereupon the crew was forced to seek shelter on Bava Island. The plane wreck lies in 3 m of water, approximately one mile due west of Bava Island (Whiting, 1994:126-27). Grid References are approximate only. B. David, B. Duncan, M. Leavesley 4-567 x Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Type of Cultural Heritage Site Comments* N Archaeological Aircraft wrecks and other WWII sites (cont’d) CB22 B-25C Mitchell Bomber – (named Draft Dodger) – Hidiha Island Remains of B-25C Mitchell Bomber Crash Site on Hidiha Reef (from Uhlig, 1943:pacificwrecks.com) On 21 February 1943, this B-25C Mitchell Bomber th rd (from the 90 Bomb Squadron, 3 Bomb Group nd piloted by 2 Lt Gordon McCoun) was undertaking bombing practice, when the bomb exploded prematurely just after being dropped. The explosion seriously damaged the plane’s fuel lines and wings, severely limiting its ability to fly. The pilots successfully crash-landed the plane on the reef-top near Hidiha (Idihi) Island, and the entire crew was subsequently rescued. Although the aircraft landed intact, it has since been severely disturbed by oceanic swells which break over the reef platform. The remains are spread over an area of 200 m along the reef in shallow water less than 1 m deep at low tide (Whiting, 1994:123-25). An A-20 Havoc bomber (from Morse, 1984:454) On 4 February 1944, this aircraft piloted by 2 Lt th th Chester Rimer (of the 386 Bomb Squadron, 312 Bomb Group) crashed into the sea between Bavo and Hidiha Island, with the loss of all its crew (Whiting, 1994:137). This plane has not been relocated. It may be archaeological site CB25, estimated to be located at the Grid References listed here (Grid References are approximate only). th CB25 A-20G Havoc (Serial # 439122) – Between Hidiha and Bava Islands B. David, B. Duncan, M. Leavesley 4-568 x nd x Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Type of Cultural Heritage Site Comments* N Archaeological Aircraft wrecks and other WWII sites (cont’d) th th P-39 Airacobra (Serial # 41 – 6945) – Redscar Bay On 13 May 1942, a P-39 Airacobra piloted by Lt Carpenter of the 35 Fighter th Squadron, 8 Fighter Group was shot down after intercepting incoming Japanese aircraft. Although several Zero aircraft were shot down, his aircraft was hit and crashed near Redscar Bay, possibly on the mudflat 25 miles northwest of Port th Moresby. The pilot parachuted and later rejoined his unit on 14 May 1942 (Air Museum, n.d:192; http://www.pacificwrecks.com/aircraft/p-39/41-6945.html). There are no reported specific site locational details for this site. P-39 F-1 Airacobra (Serial # 417136) – Last seen Redscar Bay On 16 June 1942, 2 Lt Stanley Rice (of the USAAF 5 Fighter Air Force, 35 th Flight Group, 40 Fighter Squadron) was killed in action when his P-38F-1 Airacobra was shot down after being attacked by Tainan Kokutai A6M2 Zero aircraft. The plane was last seen over Redscar Bay and Lt Rice was declared Missing in Action (http://www.pacificwrecks.com/aircraft/p-39/41-7136.html). It may be archaeological site CB24, estimated to be located at the Grid References listed here (Grid References are approximate only). B-25H-1 Mitchell Bomber (Serial # 434341) – Redscar Head-Lea Lea On 2 September 1944, 2 Lt Robert Dreger of the USAAF 5 Air Force Combat Replacement Training Center began a flight from the 7 Mile Drome airstrip (at Port Moresby) and crashed into the sea near Redscar Head. Although another plane observed an oil slick, a wrecked plane wheel and empty life rafts, there were no survivors. Lea Lea villagers reported that the aircraft was flying low, when a wing and propeller hit the water and the plane exploded (http://www.pacificwrecks.com/aircraft/b-25/43-4341.html). It may be archaeological site CB23, estimated to be located at the Grid References listed here (Grid References are approximate only). nd CB24 nd CB23 th nd B. David, B. Duncan, M. Leavesley 4-569 x th x th x Oral Tradition Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.6 List of sites previously known from Study Area or its vicinity, based on literature sources discussed in Section 4.5.8 of this report (cont’d) PNG National Museum and Art Gallery or Monash University Site Code Site Name Site Location (Grid Reference, 1:100,000 Port Moresby Map Sheet, AGD66 or WGS84) GPS points removed to protect sites – see original study (David 2008) E Comments* N Type of Cultural Heritage Site Archaeological Oral Tradition Aircraft wrecks and other WWII sites (cont’d) CB5 Telegraph Cable Telegraph Cable – Australia to Boera Point (location is unknown. Possibly comes ashore in the vicinity of Davage at the Grid References listed here) This site was not inspected, as it was only identified from historical records after fieldwork had been undertaken. Further research is required to identify its exact location, and to ensure the facility is not still being used. x * Entries in red highlight locational uncertainties. Where sites have subsequently been recorded by the cultural heritage site recording team during field surveys, site codes have been entered in the left-hand column (see also Table 4.5.8 for details of sites recorded during field surveys). B. David, B. Duncan, M. Leavesley 4-570 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.7 Site # Mean ground surface visibility by recorded indigenous archaeological site. Mean Ground visibility 0% 1-25 % 26-50 % AAHL x AAHM x AAHN 76-100 % x AAHO x AAHP x AAHQ x AAHR x AAHS x AAHT x AAHU x AAHV x AAHW x AAHX x AAHY x AAHZ x AAIB x AAIC x AAID x AAIE x AAIF x AAIG x AAIH x AAII x AAIJ x AAIK x AAIL x AAIM x AAIN AAIO 51-75 % x x AAIP x AAIQ x AAIR x AAIS x AAIT x AAIU x AAIV x LNG1 x LNG2 x LNG3 x B. David, B. Duncan, M. Leavesley 4-571 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.7 Site # LNG4 Mean ground surface visibility by recorded indigenous archaeological site (cont’d). Mean Ground visibility 0% 1-25 % ML1 ML2 26-50 % 51-75 % 76-100 % x x x ML3 x ML4 x ML5 x ML6 x ML7 x ML8 x ML9 x ML10 x ML11 x ML12 x ML13 x ML14 x ML15 x ML16 x ML17 x ML18a x ML18b x ML19 x ML20 x ML21 x JD1 x JD2 x JD3 x JD5 JD6 x JD7 x JD8 x JD9 x JD10 x JD11 x JD12 x JD13 JD14 x x JD15 x JD16 JD17 x x x = predominant degree of ground visibility. Site # refers to PNG National Museum and Art Gallery site register (site codes beginning with ‘A’) or, where these are not available, Monash University site codes (all other site codes) B. David, B. Duncan, M. Leavesley 4-572 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Effects of Ground Visibility on Site Recovery. Most of Portion 152 is covered with open lowland forest vegetation, including a complete ground cover of tall grasses and/or occasional herbaceous plants where surface visibility is minimal. Although systematic ground-walking surveys were implemented, ground visibility was considerably reduced as a result of the dense grass cover. It is significant that even where sites were found, mean ground surface visibility remained very low (<51%) in 70 (90%) of the 78 field-surveyed archaeological sites (Table 4.5.7). Of the 11 archaeological sites recorded with mean ground visibility greater than 50%, nine (82%) are beach, mangrove, river-bed or erosion-exposed deposits where the area is largely free of vegetation (sites AAIP, AAIQ, ML3, JD1, JD2, JD3, JD7, JD11 and JD13); one site (JD5) was not allocated a rating. In short, thick grass cover has meant the virtually complete covering of archaeological sites by vegetation within much of the Priority Area. This factor will need to be taken into account when recommending cultural heritage site management procedures during the construction stage of the LNG Facility. Survey Results. A total of 78 indigenous cultural heritage sites were recorded during the field surveys (42 during the 100 % surveys of January 2008, 36 during the transect surveys of AprilMay). The January sites were recorded firstly during surveys within the originally-defined kidneyshaped area, and subsequently in the updated Study Area. In all cases sites were thus located either within the Study Area or very close to it. We discuss the sites by site type, taking into account their environmental contexts. Locational and other details of each site are presented in Table 4.5.8. We point out that no sub-surface testing of any site has been attempted (as this was also beyond our brief); the following site descriptions are therefore based only on what could be seen on the site surfaces and within their exposed erosional features (giving insight into subsurface deposits). We shall return to assessments of significance, potential impacts from proposed developments, and management recommendations in Section 4.5.12, Section 4.5.13 and Section 4.5.14 below. Ethnographically-Known Villages. Four villages known from oral traditions and historical records were recorded during the present surveys (many others occur in the region, but have not been recorded as they are located some distance outside the Study Area): Konekaru (JD1+2), Darebo (CB11), Buria (CB10) and a site at Dori Hill (CB12). Archaeological site JD1+2 was identified in the field by Boera and Papa clan members as the old, historical village of Konekaru, which means in Motu ‘Coconut Beach’. Konekaru has previously been discussed from historical records in Section 4.5.6.1 of this report. Consistent with its name, Konekaru is located on the only stretch of beach in the northwestern corner of Portion 152; it currently has no evidence of coconut trees, probably indicating some antiquity (decades or more) since village abandonment. The site is an ancestral village for current Papa villagers in particular. Site CB10 is the ethnographically-known village and cultural complex at Buria Hill (see Plate 4.5.26), which was once the site of a village near Lea Lea populated by the Namura and Arauwa groups. Several versions of the destruction of the village exist, and comment on this village has already been made in relation to the literature review in Section 4.5.10.1.1 above. One version recounts how, after a Buria chief mistreated his wife from Badihgwa and robbed her family’s lagatoi as it passed his village, his wife schemed to bring about his downfall. One night when the villagers were unarmed during a dancing performance, the wife’s relatives attacked and slaughtered all but one person in the village. An alternative version recounts how the Buria chief rejected a Pari woman betrothed to him. In retaliation for the insult, the Pari people destroyed the village during a dance ceremony, leaving only one woman alive. The incident was supposed to have occurred ten generations ago (Kanasa, 2006:13) (see Section 4.5.10.1.1 for further details of oral traditions). Lea Lea villagers spoke in reverential terms of Buria as a traditional place. All of those we have spoken with indicated that this is a site where today people show deep respect and speak in B. David, B. Duncan, M. Leavesley 4-573 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project hushed tones. Councillor Hene Totana (personal communication 2008) indicated that there were ceramic sherds ‘everywhere’ at this traditional site, along with wooden tools (for digging gardens) which still stand exactly where the people had left them. Some of these tools have been burnt during recent fires. The Lea Lea people are very keen to have this site documented, as they are adamant that that it must be protected. Plate 4.5.26 Buria Hill (viewed from the south) The traditional ancient village site of Darebo (CB11) is located on Darebo Hill to the northeast of Buria, and is shown on the map of traditional names by Auda Delena. It was identified as a significant cultural site by community representatives, but its specific locational and cultural characteristics and significance remain to be recorded. Because it lies some distance from the proposed developments, we mention this site here without further details. There is an ancient village site on Dori Hill (CB12); although it is not shown on the traditional names map by Auda Delena, it is the hill located to the adjacent east of Buria. No further information was supplied by community representatives for this site. Local villagers also mentioned another ancestral village on a hilltop in the southeastern corner of Portion 152. However, as it lay outside the Study Area of the brief and due to time constraints, this site was not able to be recorded for the present report. Human Burials/Cemeteries. No evidence of human burials or cemeteries has been identified in Portion 152. However, PNG ethnography suggests that in previous generations the custom of burying the dead under existing houses was prevalent. As one ethnographically known village site (Konekaru) has been identified within Portion 152, and as every archaeologically excavated village site whose cultural materials have been reported in the Port Moresby region has revealed human remains, the likelihood of future discoveries of human burials or human remains in Portion 152 is high. Indeed it is likely that close relatives of present-day villagers, such as grandparents and great grandparents, are buried at Konekaru and/or other nearby sites. Furthermore, in addition to the ethnographic village of Konekaru, other large sites and site complexes recorded during our field surveys (see below and Table 4.5.8 for details) are interpreted by us as archaeological evidence of old villages not identified as such in the oral traditions we have recorded (possibly because those archaeological villages are too old to have been remembered by present-day communities). As ancient village sites, each is likely to contain human remains. This issue will be considered in Section 4.5.14 when discussing the management of the area’s cultural heritage places. Sacred Sites and Traditional Story Sites. Community representatives from Lea Lea and Papa told members of the cultural heritage site survey team that, additional to the ancestral village sites discussed above, there are a number of sacred sites in the hills to the east of Portion 152. Sacred B. David, B. Duncan, M. Leavesley 4-574 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project sites can occur in any environment, including trees, rock outcrops, waterways, hills and plains. These sites involve various kinds of spirits and ancestors associated with local villagers. As this area is outside Portion 152 (and outside our brief), we have not visited nor recorded these sites. However, given the anticipated population increases and cumulative demographic impacts to result from the planned LNG152 Facility developments, such nearby cultural heritage sites will almost certainly be impacted through visitation and the like (e.g., local people opportunistically expanding their own ventures, e.g., gardening, shifts in camping or housing locations). This point will be revisited in Section 4.5.14 below. Some of the villages documented through oral traditions – such as Buria, discussed above – can also be thought of as sacred sites, due to the ancestral spirits that reside there. Such information has not been recorded for the present report, as these sites lie outside the Study Area and will require specialist attention in the future. Another kind of oral tradition site of great cultural significance that has been recorded are places associated with the Edai Siabo first lagatoi ancestral story. Two such sites have been recorded on land during our surveys: CB7 and CB8. Two other story sites (CB6 and ASM) also relating to the Edai Siabo story occur in the shallow marine environment and are thus presented separately in Section 4.5.10.2.2 below. CB7 was identified by Daro Avei, a fisherman from Boera village. It is located between Boera and Vaihua Creek. Avei (personal communication 2008) maintains that stone flakes produced when making the stone axes to carve the first canoe are evident during the dry season in this area. This is a traditional cultural site where the wood was felled and roughly shaped before the canoe was transported through the mangroves to the ancient village where Edai Siabo lived near Davage. This was the only instance during our surveys that we heard the story of this site being told. As far as we know, this is the first time that this story place has been recorded. It is probable that this site falls within the bounds of the Portion 152 area (but it lies outside the Study Area), but its exact location could not be visited as the informant had other commitments and the area was off-limits to survey staff due to the UXO discovery. Further work is required to determine the exact location and significance of this site. Plate 4.5.27 Left: Stream outlet at Davage, where Edai Siabo’s first lagatoi was built. Plate 4.5.28 Right: Dugout canoe under construction at Davage. CB8 was identified as the location where the first lagatoi was built, and is situated in the first small bay to the north of Boera Head (to the south of the Marine Offloading Facility area). It was located close to a stream outlet which collects into a small pool just above the high tide mark on the northern extremity of the bay (known locally as Davage), where fishermen were said to have washed after they returned from their day at sea fishing (Plate 4.5.27). The site is currently being B. David, B. Duncan, M. Leavesley 4-575 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project used to build a dugout canoe (Plate 4.5.28), evident in the scatters of woodchips relating to its construction. There is also a modern timber platform which has been built just above the beach under a large tree to be used as a place to rest in the shade. The ancient village of Apau, which predates the current village of Boera, was located close-by approximately 70 m away on the southern edge of this bay, but was not relocated during our brief visit due to high vegetation cover. Wooden posts and ceramics from this former stilt village were reported to still be evident during the dry season (Moi Dobi, personal communication 2008). Archaeological Scatters of Pottery, Stone Artifacts and Midden Shell. The most prevalent archaeological site type in the surveyed area contained pottery, stone artifacts and midden shell material. A total of 22 such sites were recorded during the January surveys. There is a concentration along a strip of land that follows a broad north-south trajectory between the eastern shoreline of the mudflats that is the southern arm of the tidal inlet and the Ruisasi Creek. These sites consist of: AAHM, AAIT, ML1, ML2, ML3, ML4, ML8, ML10, ML15, ML16, ML18, ML19, ML20, JD5, JD6, JD10, JD11, JD12, JD13, JD15, JD16, and JD17. The largest of these are in the vicinity of the ethnographic village of Konekaru (JD1+2) to the north and JD8 to the south. Most of the other, smaller sites of this type are scattered in-between. One site (JD13) stands out from the others because obsidian flakes were identified within the stone artifact assemblage (see 4.5.12 for the significance of obsidian artifacts; see 4.5.8.3 and 4.5.8.7 above for discussions of obsidian in southern PNG coastline archaeological sites). Archaeological Scatters of Pottery Only. Five such sites were recorded: AAIG, LNG1, LNG2, ML6 and JD9. One site (ML6) is located at the northern end of a low ridge adjacent to the mudflats that parallel the southern arm of the tidal inlet. Three sites (AAIG, LNG1 and LNG2) are located on grassy low-gradient hill-slopes immediately west of the Boera-Papa Road. The remaining site (JD9) is located north of and in-between the southern arm of the tidal inlet and Ruisasi Creek. Archaeological Scatters of Pottery and Stone Artifacts Only. Eight sites of this type were recorded: AAHU, AAIR, LNG3, LNG4, ML5, ML9, ML12 and ML14. Four of these (LNG3, LNG4, ML5 and ML9) are located on the grassy low-gradient hill-slopes immediately west of the BoeraPapa Road; AAHU is located in a similar setting, to the east of that road. The remaining sites (AAIR, ML12 and ML14) are located between the southern arm of the tidal inlet and Ruisasi Creek. Archaeological Scatters of Pottery and Midden Shell Only. Eleven such sites were recorded: AAIB, AAIC, AAIN, AAIO, AAIU, ML11, ML13, ML17, ML21, JD1+2 and JD7. Two sites (ML11 and JD7) are located along the low ridge adjacent to the mudflats at the southern arm of the tidal inlet. AAIO is also found on the edge of mudflats just to the west of the Boera-Papa road. AAIN is located among mangroves. One site (JD1+2) is spatially close and in line with the old village site at Konekaru. AAIB is found on grassland towards the top of a hillslope to the west of the Boera-Papa road, AAIC and AAIU to the east. The remaining sites (ML13, ML17 and ML21) are located along the western bank of Ruisasi Creek. Archaeological Scatters of Stone Artifact Only. Seven such sites were identified: AAHL, AAHX, AAIM, AAIP, AAIQ, ML1 and ML7. AAHL and AAHX are both located >1.5km from the coast on low, grassy ridge-slopes. AAIM is found on the edge of flood plains, where grasslands meet the clay pan; AAIQ near the edge of mudflats and grassland; and AAIP on mudflats. ML1 and ML7 are located in relatively close proximity to Konekaru beach. Archaeological Scatters of Midden Shell Only. Seven such sites were recorded. They are: AAHP, AAHR, AAHV, AAID, AAII, JD3 and JD14. AAHP, AAHR, AAHV, AAID and AAII are small, low-density shell scatters mostly located >1 km inland on low grassy hill-slopes or, in the case of AAID, in similar settings on grassy plains near a small ephemeral creek. JD3 is extensive; it is located near and along the same sand dune that connects with Konekaru beach. This dune B. David, B. Duncan, M. Leavesley 4-576 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project extends northwards behind the fringing mangroves. JD3 is located very near the boundary of Portion 152. The remaining site (JD14) is located on the grassy low ridge that runs parallel to the mudflats at the southern tidal outlet draining towards Konekaru beach. Archaeological Scatters of Pottery, Stone Artifacts, Midden Shell and Animal Bones. Only two such sites, AAHN and JD8, were recorded. The absence of vertebrate faunal remains has been a feature of most recorded sites, indicating a focus on the marine environment (evidenced in the shell remains), as typical of ethnographically documented Motu coastal sites. AAHN is an extremely rich archaeological site, ~1.0 to 1.5 km inland at the base of a low ridge, at an ecotone between the grassland and clayey floodplain. It probably represents a past settlement. JD8 is located along the back dune towards the northwestern end of the Priority Area. It is near Konekaru beach, and may have been functionally related to the old settlement of Konekaru. Archaeological Scatters of Stone Artifacts and Midden Shell Only. Three such sites have been recorded: AAHO, AAHS and AAIJ. These are each located >1 km from the coast, and are predominantly small shell middens with a few stone artifacts. Archaeological Sites Containing Isolated Stone Artifacts Only. Two sites (AAHT, AAHW) consist of isolated stone artifacts exposed on the ground surface. They are each located on low grassy ridgeslopes >1.5 km inland. Archaeological Sites Containing Isolated Pottery Sherds Only. Two sites (AAHY and AAIS) consist of isolated pottery sherds exposed on the ground surface. They were both found on low, grassy rigde-slopes. Archaeological Sites Containing Isolated Shells Only. Eight sites (AAHQ, AAHZ, AAIE, AAIF, AAIH, AAIK, AAIL, AAIV) consist of isolated shells exposed on the ground surface. They were each found on low, grassy rigde-slopes, at the base of a ridge or on plains >1.5 km inland. Early Missionary Sites. A memorial commemorating the first landing of Reverend William Lawes at Boera is located on the spot where he came ashore and the first church was subsequently constructed in the village (Douglas, 2007) (see Plate 4.5.29). This site has been recorded as CB19. The London Missionary Society (LMS) has played a large role in village life at Boera, Papa and Lea Lea, all of which include significant LMS churches. A large tree on the banks of Lea Lea Creek inside Lea Lea village is considered locally significant for its associations with the London Missionary Society (see Plate 4.5.30). It was planted by the first missionary as a gesture to the community when he arrived at the village (Igo Meauri, personal communication 2008). The tree is in danger of being washed away or being affected by saltwater inundation if coastal erosion in this area continues. It is recorded recorded as cultural heritage site CB20. B. David, B. Duncan, M. Leavesley 4-577 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Site Must Not Be Damaged Recommend Stage 2 Archaeological Salvage WWII Plane Crash WWII Defence Facility WWII Shipwreck Lagatoi Wreck Tree Spirit/Sacred/Story Place Early Missionary Site Old Village (from Oral Traditions) Metal (Includes Bullets) Glass Animal Bones Stone Artifacts Shell Pottery Grid Reference (deleted to protect site location – see original report, David 2008) UTM Date of Site Recording Site Name Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys PNG Museum and Art Gallery Site Code Monash University Site Code Table 4.5.8 Comments LNG1 N/A Konekaru (East) 17-Jan-08 55S x x This is an area (16 x 1 m) of low-density archaeological material including pottery (<10 sherds) only. Site is located on the west side of the Boera/Papa Rd. Situated on a low-gradient grassy hill-slope. Associated story: approximately 300 m north is a ‘nose-piercing place’. From this place a man could see through the hole in another man’s nose and see another person working in a garden – ‘Urivaka’. It is considered part of the Konekaru region, although the central Konekaru location is ~1 km to the west. LNG2 N/A Konekaru (East) 17-Jan-08 55S x x This is an area (12 x 6 m) of low-density archaeological material including pottery (4 sherds) only. Site is located on the west side of the Boera/Papa Rd (west of LNG1). Situated on a low-gradient grassy hill-slope. LNG3 N/A Konekaru (East) 17-Jan-08 55S x x x This is an area (8 x 19 m) of low-density archaeological material including pottery (8 sherds) and a single stone artifact (chert core). Site is located west of Boera/Papa Rd (west of LNG2). Situated on a low-gradient grassy hill-slope. LNG4 N/A N/A 17-Jan-08 55S x x x This is an area (4 x 5 m) of low-density archaeological material including pottery (1 sherd) and 3 stone artifacts (silcrete/chert). Site is located on the west side of the Boera-Papa Rd (west of LNG3). Situated on a low-gradient grassy hill-slope. ML1 N/A N/A 18-Jan-08 55S x x This is an area (15 x 7 m) of low-density archaeological material including 3 angular fragments of red chert. Site is located on the west side of the Boera-Papa Rd (west of LNG4). Situated on a very low-gradient grassy floodplain that drains west to the tidal inlet of which Konekaru is adjacent. ML2 N/A N/A 18-Jan-08 55S x x This is an area (35 x 10 m) of low-density archaeological material including pottery (11-50 sherds), stone artifacts (N = 11-50) and midden shell. Site is located on the eastern edge of the tidal inlet associated with Konekaru. Situated on a very low-gradient floodplain grassland. x This is an area (20 x 1 m) of high-density archaeological material including pottery (11-50 frags), stone artifacts (N = 1-10) and midden shell (N = 150+). Site is located on the northern margin of the tidal inlet associated with Konekaru and is probably part of the main Konekaru site complex. Situated on the eroding Section of the northern bank of the tidal inlet. x x This is an area (4 x 5 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10), and marine shell (N = 1-10). Site is located west of the Boera-Papa Rd (south of ML1). It is probably the southern extension of ML1. Situated on a low-gradient grassy hill-slope. x x This is an area (50 x 40 m) of low-density archaeological material including pottery (1-10 sherds) and stone artifacts (N = 1-10). Site is located west of the Boera-Papa Rd (south of ML4). It is probably the southern extension of ML4. Situated on a low-gradient grassy hill-slope. x This is an area (15 x 10 m) of low-density archaeological material including pottery (1-10 sherds). Site is located on the eastern bank of the southern tidal inlet extending south from Konekaru. It is situated on the northern end of a low ridge that runs south parallel to the tidal inlet and ultimately the coast. x This is an area (70 x 25 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 1-10). Site is located west of the Boera-Papa Rd (south of ML5). It is probably the southern extension of ML5. Situated on a low-gradient grassy hill-slope. x x ML3 N/A Konekaru 18-Jan-08 55S x x ML4 N/A Konekaru 18-Jan-08 55S x x ML5 N/A N/A 18-Jan-08 55S x ML6 N/A N/A 18-Jan-08 55S x ML7 N/A N/A 18-Jan-08 55S x x x B. David, B. Duncan, M. Leavesley 4-578 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project ML8 N/A N/A 18-Jan-08 55S x ML9 N/A N/A 18-Jan-08 55S x ML10 ML11 N/A N/A N/A N/A 19-Jan-08 19-Jan-08 55S x 55S N/A N/A 19-Jan-08 55S x ML13 N/A N/A 19-Jan-08 55S x ML15 ML16 N/A N/A N/A N/A N/A N/A 19-Jan-08 19-Jan-08 19-Jan-08 55S 55S 55S x x x x x x This is an area (30 x 20 m) of low-density archaeological material including pottery (10 sherds) and stone artifacts (N = 1-10). Site is located west of the Boera-Papa Rd (south of ML5). It is probably the southern extension of ML4, ML5 and ML7. Situated on a low-gradient grassy hill-slope. x This is an area (100 x 20 m) of low-density archaeological material including pottery (11-50 sherds), stone artifacts (N = 11-50) and midden shell (N = 1-10). Site is located on the eastern bank of the southern arm of the tidal inlet and is probably the southern extension of ML6 and ML8. It is situated on a low ridge that runs south parallel to the tidal inlet and ultimately the coast. x This is an area (5 x 5 m) of low-density archaeological material including pottery (25 sherds) and midden shell (N = 1-10). Site is located on the eastern bank of the southern tidal inlet and is probably the southern extension of ML6, ML8 and ML10. It is situated on a low ridge that runs south parallel to the tidal inlet and ultimately the coast. x This is an area (20 x 20 m) of low-density archaeological material including pottery (~10 sherds, including 1 decorated rim sherd) and stone artifacts (N = 1-10 quartz). Site is located on the western bank of the Ruisasi Ck. It is southwest of ML13. The site is situated at the base of a lunette that marks the western edge of the floodplain associated with Ruisasi Ck. x This is an area (5 x 25 m) of low-density archaeological material including pottery (1 sherd) and midden shell (N = 11-50). Site is located in the Ruisasi Ck bed and is most likely to be a secondary deposit. There is no evidence of nearby in situ deposits. The site is northwest (upstream) of ML14. x x x x Site Must Not Be Damaged Recommend Stage 2 Archaeological Salvage WWII Plane Crash WWII Defence Facility WWII Shipwreck Lagatoi Wreck Tree Spirit/Sacred/Story Place Early Missionary Site Old Village (from Oral Traditions) Metal (Includes Bullets) Glass Animal Bones Stone Artifacts x x x Comments This is an area (30 x 10 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 1-10). Site is located on the eastern bank of the southern tidal inlet extending south from Konekaru and is probably the southern extension of ML6. It is situated on a low ridge that runs south parallel to the tidal inlet and ultimately the coast. x x ML12 ML14 Shell Pottery Grid Reference (deleted to protect site location – see original report, David 2008) UTM Date of Site Recording Site Name Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys (cont’d) PNG Museum and Art Gallery Site Code Monash University Site Code Table 4.5.8 x x B. David, B. Duncan, M. Leavesley 4-579 x This is an area (20 x 5 m) of medium-density archaeological material including pottery (11-50 sherds) and stone artifacts (N = 1-10). Site located on a lunette east of ML13 where the creek changes its course from west to south. The lunette appears to be a remnant from the periodic flooding of the Ruisasi Ck. The site has potential for sub-surface deposit. x This is an area (40 x 15 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 1-10). Site located on the grassy higher ground that runs N-S on the eastern side of the mudflat that is the southern arm of the tidal inlet that drains to Konekaru. The site is south of ML11 and north of ML16. Evidence from other parts of the ridge suggests that this site may have sub-surface deposits. x This is an area of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 110, including a stone club-head) and midden shell (N = 1-10). Site located on the grassy higher ground that runs N-S on the eastern side of the mudflat that is the southern arm of the tidal inlet that drains north to Konekaru. The site is south of ML15 and north of ML19. Evidence from other parts of the ridge suggests that this site may have sub-surface deposits. Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project ML18 ML19 N/A N/A N/A N/A 19-Jan-08 19-Jan-08 55S 55S x x x x ML20 N/A N/A 20-Jan-08 55S x x ML21 N/A N/A 20-Jan-08 55S x x JD1+2 N/A Konekaru 18-Jan-08 55S JD3 N/A Konekaru 18-Jan-08 55S JD5 N/A North Konekaru 18-Jan-08 55S x x x x x x x x x x B. David, B. Duncan, M. Leavesley 4-580 Site Must Not Be Damaged Recommend Stage 2 Archaeological Salvage WWII Plane Crash WWII Defence Facility WWII Shipwreck Lagatoi Wreck Tree Spirit/Sacred/Story Place Early Missionary Site Old Village (from Oral Traditions) x Metal (Includes Bullets) x Glass 55S Animal Bones 19-Jan-08 Stone Artifacts UTM N/A Shell Date of Site Recording N/A Pottery Site Name ML17 Grid Reference (deleted to protect site location – see original report, David 2008) PNG Museum and Art Gallery Site Code Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys (cont’d) Monash University Site Code Table 4.5.8 Comments x This is an area (7 x 7 m) of low-density archaeological material including pottery (1-10 sherds) and midden shell (N = 1-10). Site located on a lunette that marks the western extent of the periodic flooding of the Ruisasi Ck. The site is south of ML14 and north of ML20. Site has little evidence of sub-surface deposits. x ML18a is an area (20 x 20 m) and ML18b (7 x 7m) of low-density archaeological material including pottery (11-50 sherds), stone artifacts (N = 1-10, including obsidian) and midden shell (N = 11-50). Sites are located on the floodplain on the western bank of the Ruisasi Ck, south of ML17 and north of ML20. Sites have little evidence of sub-surface deposit. These sites were recorded together because of their spatial proximity – they are potentially part of a single site. x This is an area (35 x 20 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 1-10). Site located on the grassy higher ground that runs N-S on the eastern side of the mudflat that is the southern arm of the tidal inlet that drains north to Konekaru. The site is south of ML16. Evidence from other parts of the ridge suggests that this site may have sub-surface deposits. x This is an area (25 x 15 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 1-10). ). Site located on the southern (of two) lunette(s) that marks the western extent of the periodic flooding of the Ruisasi Ck. The site is south of ML17 and immediately north of ML21. Site has little evidence of sub-surface deposits. x This is an area (35 x 20 m) of medium-density archaeological material including pottery (11-50 sherds) and midden shell (N = 1-10). Site is located clearly in the floodplain that is the western bank of the Ruisasi Ck. It is south of ML20. Irrespective of the geographical context it has potential for sub-surface deposits. x This is an area (148 x 15 m) of high-density archaeological material including pottery (50+ sherds) and midden shell. Site is located on the Konekaru beach. Currently the only beach within Portion 152. The associated story describes how Konekaru was a former village site before the inhabitants moved to Papa. The story places the site well within living memory. The site also contains evidence of more recent use including fencing and used rifle cartridges. From an archaeological perspective the site may be connected with JD3 and JD5. Surface indications suggest that this site is highly likely to have a sub-surface component. x x This is an area (4 x 1.5 m) of medium-density archaeological material including midden shell (N = 50+). Site is located on the beach sand behind the mangroves north of Konekaru at the northern extension of Portion 152. Surface indications suggest that this site is highly likely to have a sub-surface component. x This is an area (200 x 40 m) of high-density archaeological material including pottery (50+ sherds), stone artifacts (N = 11-50) and midden shell (N = 50+). Site is located 100 m east of Konekaru beach. It is eroding out of the higher ground that forms a peninsula between the southern arm of the tidal inlet and the eastern arm. From an archaeological perspective it is highly likely that it is associated with the Konekaru beach site (JD1+2). Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project JD7 N/A N/A 19-Jan-08 55S x JD8 N/A N/A 19-Jan-08 55S x JD9 N/A N/A 19-Jan-08 55S x JD10 N/A N/A 19-Jan-08 55S x x x x x x x x Site Must Not Be Damaged Recommend Stage 2 Archaeological Salvage WWII Plane Crash WWII Defence Facility WWII Shipwreck Lagatoi Wreck Tree Spirit/Sacred/Story Place Early Missionary Site Old Village (from Oral Traditions) x Metal (Includes Bullets) x Glass x Animal Bones 55S Stone Artifacts 19-Jan-08 Shell UTM N/A Pottery Date of Site Recording N/A Grid Reference (deleted to protect site location – see original report, David 2008) Site Name JD6 Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys (cont’d) PNG Museum and Art Gallery Site Code Monash University Site Code Table 4.5.8 Comments x This is an area (20 x 13 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 50+). Site located on the higher ground east of the southern arm of the tidal inlet. Surface indications suggest that this site is highly likely to have a sub-surface component. x This is an area (50 x 35 m) of low-density archaeological material including pottery (1-10 sherds) and midden shell (N = 50+). Site is located in the transitional zone between the mudflats of the tidal inlet and the grassland on the higher ground. Surface indications suggest that this site is highly likely to have a subsurface component. This site is likely to be an extension of JD6. x This is an area (50 x 20 m) of high-density archaeological material including pottery (11-50 sherds), stone artifacts (N = 1-10), midden shell (N = 50+) and animal bone (N = 1-10). Site located at the confluence of the Ruisasi Ck and the mudflats that are the tidal inlet. Surface indications suggest that this site is highly likely to have a sub-surface component. This site is likely to be an extension of JD5 and JD7. x This is a small area (1 x 1 m) of high-density archaeological material including pottery (1-10 sherds). Site is located on the raised grassy ground that is the west bank of the Ruisasi Ck between JD10 to the north and JD8 to the south. Surface indications suggest that this site is highly likely to have a sub-surface component. x This is an area (50 x 50 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 1-10). Site is located on the raised grassy ground that is the west bank of the Ruisasi Ck between JD9 and JD11. Surface indications suggest that this site is highly likely to have a sub-surface component. The site is close to JD9 and may be and extension of it. JD11 N/A N/A 19-Jan-08 55S x x x x This is an area (5 x 65 m) of high-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 11-50). Site is located on the raised grassy ground that is the north (nominal western side), bank of the Ruisasi Ck between JD10 to the south and JD13 to the north. Surface indications suggest that this site is highly likely to have a sub-surface component. JD12 N/A N/A 19-Jan-08 55S x x x x This is an area (5 x 21 m) of high-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 50+). Site is located ~70 m north of JD11 on a lunette adjacent to the Ruisasi Ck. Surface indications suggest that this site is highly likely to have a sub-surface component. x This is an area (40 x 102 m) of high-density archaeological material including pottery (11-50 sherds), stone artifacts (N = 1-10) including obsidian and midden shell (N = 50+). Site located in the bed of the Ruisasi Ck. Origin of the archaeological material is thought to be the nearby northwest creek bank. The site is located between JD12 to the south and JD14 to the north. Visual inspection of creek bank suggests high potential for sub-surface deposits. x This is an area (40 x 60 m) of medium-density archaeological material including shell (N = 50+). Site located in the bed of the Ruisasi Ck. Origin of the archaeological material is thought to be the nearby northwest creek bank, although there is no evidence of associated sub-surface deposits. The site is located between JD13 to the south and JD15 to the north. JD13 JD14 N/A N/A N/A N/A 19-Jan-08 19-Jan-08 55S 55S x x x x B. David, B. Duncan, M. Leavesley 4-581 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project JD15 JD16 N/A N/A N/A N/A N/A 19-Jan-08 20-Jan-08 20-Jan-08 55S 55S 55S x x x x x x x x JD17 N/A x JDA1 AAHL 29-Apr-08 JDA2 AAHM 29-Apr-08 55S x x x JDA3 AAHN 29-Apr-08 55S x x x JDA5 AAHO 29-Apr-08 55S x x JDA6 AAHP 30-Apr-08 55S x JDA7 AAHQ 30-Apr-08 55S x JDA8 AAHR 30-Apr-08 55S x JDA9 AAHS 30-Apr-08 55S x JDA10 AAHT 30-Apr-08 55l JDA11 AAHU 30-Apr-08 55S JDA12 AAHV 30-Apr-08 55S JDA13 AAHW 30-Apr-08 55S x JDA14 AAHX 30-Apr-08 55S x Comments x This is an area (19 x 4 m) of high-density archaeological material including pottery (50+ sherds), stone artifacts (N = 11-50) and midden shell (N = 50+). Site is located on the high grassy ground that is the west bank of the Ruisasi Ck, between JD14 to the south and JD16 to the north. Surface indications suggest that this site is highly likely to have a sub-surface component. This is an extremely rich archaeological deposit. x This is an area (160 x 40 m) of medium-density archaeological material including pottery (11-50 sherds), stone artifacts (N = 1-10) and midden shell (N = 50+). Site is located in the bed of the Ruisasi Ck north of JD15. Archaeological material is thought to have eroded from the nearby northwest creek bank; no in situ material was identified. There is some suggestion that the material may be eroding from the bed of the creek itself. A drilled disk was collected from the site under a permit to Mr John Dop from the PNG National Museum and Art Gallery. x This is an area (13 x 14 m) of low-density archaeological material including pottery (1-10 sherds), stone artifacts (N = 1-10) and midden shell (N = 11-50). Site is located ~30 m east of the mudflat of the tidal inlet on the grassy higher ground between JD6 to the south and JD7 to the north. Evidence associated with nearby sites suggests that this site also has potential for sub-surface deposits. A 12 x 7 m scatter of <10 stone artifacts, with a small number of shells located approximately 3 m from the stone artifacts. x x x An archaeological site extending 300 x 10 m along low-lying area near Ruisasi Ck. Consists of a light scatter of midden shell (N = 11-50), stone artifacts (N = 1-10) a pottery (1-10 sherds). x A site 70 x 15 m in area, located at ecotone of grassland and Ruisasi Ck clayey flood-plain. Very rich cultural heritage site containing 1-10 stone artifacts, 50+ shells, 50+ pottery sherds and 1-10 animal bones. x This is an area some 11 m long containing 1 stone artifact and 11-50 midden shells, among grassland. x This 8 x 6 m archaeological site is located in low undulating grassland, towards the top of a gentle slope. This scatter of cultural materials is of low density, and contains 1-10 midden shells. This site consists of a single shell on the top of a low grassland slope. x x This site is 25 m long and consists of low density midden shell (N = 1-10 shells) on low undulating grassland. x This site is 25 m long and consists of low density midden shell (N = 11-50) and stone artifacts (N = 1-10) on a gentle slope among grassland. x x Site Must Not Be Damaged Recommend Stage 2 Archaeological Salvage WWII Plane Crash WWII Defence Facility WWII Shipwreck Lagatoi Wreck Tree Spirit/Sacred/Story Place Early Missionary Site Old Village (from Oral Traditions) Metal (Includes Bullets) Glass Animal Bones Stone Artifacts Shell Pottery Grid Reference (deleted to protect site location – see original report, David 2008) UTM Date of Site Recording Site Name Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys (cont’d) PNG Museum and Art Gallery Site Code Monash University Site Code Table 4.5.8 This site consists of a single chert core on a gentle grassy slope. x x x Scatter of 2 pottery sherds and 1 chert flake spread over an area 1 x 1 m in grassland. x 2 shells spread over an area 5 x 1 m in grassland. A single chert core located near the base of a gentle grassy slope. x B. David, B. Duncan, M. Leavesley 4-582 2 flakes and 1 pottery sherd on a gentle grassy slope. Scatter spread over an area 4 x 4 m. Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project JDA17 Site Must Not Be Damaged Recommend Stage 2 Archaeological Salvage WWII Plane Crash WWII Defence Facility WWII Shipwreck Lagatoi Wreck Tree Spirit/Sacred/Story Place Early Missionary Site Old Village (from Oral Traditions) 55S Metal (Includes Bullets) 1-May-08 Glass AAHZ Animal Bones JDA16 Stone Artifacts 55S Shell UTM 1-May-08 Pottery Date of Site Recording AAHY Site Name JDA15 Grid Reference (deleted to protect site location – see original report, David 2008) Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys (cont’d) PNG Museum and Art Gallery Site Code Monash University Site Code Table 4.5.8 x Comments 1 pottery sherd on gentle grassy slope. x 1 shell on gentle grassy slope. x x x Site of a probably WWII plane crash. A local Boera man showed the surveying team this site. Local villagers recount this as the site of a WWII plane crash. Scattered remains of fuselage still present. However, much of the wreck has already been recovered by local villagers for scrap metal. A sweep of the metal detector reveals concentrations of aluminium sheeting in this area. Site spread over area of 30 x 30 m on gentle grassy slope. 1-May-08 55S 3-May-08 55S x x x 1 shell and 1 pottery sherd on gentle grassy slope. 55S x x x 1 shell and 1 pottery sherd on gentle grassy slope. JDA18 AAIB JDA19 AAIC JDA20 AAID JDA21 JDA22 MLA1 AAIG 29-Apr-08 55S MLA2 AAIH 29-Apr-08 55S x 1 shell on gentle grassy slope. MLA3 AAII 29-Apr-08 55S x 2 shells on gentle grassy slope spread over 5 m radius on gentle grassy slope. MLA4 AAIJ 30-Apr-08 55S x MLA5 AAIK 30-Apr-08 55S x 1 shell on gentle grassy slope. MLA6 AAIL 30-Apr-08 55S x 1 shell on gentle grassy plain. MLA7 AAIM 1-May-08 55S MLA8 AAIN 1-May-08 55S x MLA9 AAIO 2-May-08 55S x MLA10 AAIP 2-May-08 55S MLA11 AAIQ 3-May-08 55S MLA12 AAIR 3-May-08 55S x MLA13 AAIS 5-May-08 55S x MLA14 AAIT 8-May-08 55S x 8-May-08 55S x 2 shells on grassy plain near a small ephemeral creek, spread over 15 m. AAIE 55S x 1 shell in grassland 10 m to south of creek. AAIF 55S x 1 shell at base of gentle grassy slope. x x x x 2 conjoining pottery rim sherds spread over 5 x 5 m on gentle grassy slope. Small scatter of midden shell and a chert flake exposed through erosion among brackish swamp with 2 grassland. Patches of mangrove and Pandanus present nearby. Site spread over an area 5 m . x Small stone artifact scatter (chert) eroding from ground at ecotone between grassland and clay pan. Site 2 spread over an area 10 m . x x Scatter of shells and pottery sherds spread over 20 x 10 m on sandy substrate in mangroves. Site subject to inundation at high tide. x x Scatter of shell and pottery sherds spread over 15 x 15 m on edge of mudflats and grassland. Cultural materials are eroding out into the mudflats. Rim sherds are present. x x Scatter of 50+ chert artifacts spread over 70 x 70 m among mangroves. x x Scatter of 1000+ chert artifacts spread over 300 x 10 m among mangroves and grassland. x x Concentration of pottery sherds (N = <100) and chert flakes spread over 2 x 2 m among gentle grassy slope along the southern side of Ruisasi Ck. x 1 pottery sherd on gentle grassy slope. x x x B. David, B. Duncan, M. Leavesley 4-583 Scatter of midden shell, pottery sherds, chert flakes and a probably earth oven cobble on gentle grassy slope on side of a small knoll. Site spread over 45 x 3 m. Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Site Must Not Be Damaged Recommend Stage 2 Archaeological Salvage WWII Plane Crash WWII Defence Facility WWII Shipwreck Lagatoi Wreck Tree Spirit/Sacred/Story Place Early Missionary Site Old Village (from Oral Traditions) Metal (Includes Bullets) Glass x Animal Bones 55S x Stone Artifacts 8-May-08 Shell 55S Pottery AAIV 8-May-08 Grid Reference (deleted to protect site location – see original report, David 2008) MLA16 UTM AAIU Date of Site Recording MLA15 Site Name PNG Museum and Art Gallery Site Code Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys (cont’d) Monash University Site Code Table 4.5.8 x Comments Low concentration of pottery sherds and midden shells spread over 7 x 3 m on gentle grassy slope. x 1 shell on gentle grassy slope. CB10 Buria Hill 55S x x x This is a sacred site and ancestral village of oral tradition. It needs to be systematically recorded in the field. CB11 Darebo Hill 55S ? ? ? ? ? ? ? ? x ? ? ? ? ? x This important oral tradition site occurs some distance from the proposed LNG Facilities site. Approximate WGS84 location only. CB12 Dori Hill 55S ? ? ? ? ? ? ? ? x ? ? ? ? ? x This important oral tradition site occurs some distance from the proposed LNG Facilities site. Approximate WGS84 location only. CB7 First Lagatoi Hull Tree Felling site 55S x x This is an important site of oral tradition just north of Boera. CB8 First Lagatoi Building site (Davage) 55S x x This is an important site of oral tradition at Davage just north of Boera. CB19 Boera First Church Memorial 55S x x This is the site of the first missionary landing. It is an important historical site. CB20 Lea Lea Missionary Tree 55S x x This is an important historical site relating to the early missionary period. Apau village 55S CB30 Boera Battery 55S CB6 First Lagatoi Story Sea Cave 55S x This is an ancestral village site near Boera. Approximate WGS84 location. Site needs to be systematically recorded in the field. x x x B. David, B. Duncan, M. Leavesley 4-584 x This is an important WWII facility site. It needs to be systematically recorded in the field. x This is an important site of oral tradition near Hidiha Island. Approximate WGS84 location. Site needs to be systematically recorded in the field. Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Recommend Stage 2 Archaeological Salvage CB1 Konekaru Beach 55S x CB3 Pipe Bridge 55S CB16 Sunken Lagatoi 1 55S x x Sunken lagatoi off Lea Lea beach (near Atuha Iduka) (approximate WGS84 Location). Site needs to be systematically recorded in the field. CB17 Sunken Lagatoi 2 55S x x Sunken lagatoi southwest of Papa (exact location unknown). Site needs to be systematically recorded in the field. CB18 Sunken Lagatoi 3 55S x x Sunken lagatoi at Kerama (exact location unknown). Site needs to be systematically recorded in the field. CB21 Japanese Zero Plane Crash Site 55S x x Aircraft wreck site near Lea Lea (Kirima Iduka) (approximate WGS84 location). Site needs to be systematically recorded in the field. CB22 Draft Dodger B25C Mitchell Bomber Crash Site 55S x x Aircraft wreck site at Hidiha Island. CB23 55S x x Aircraft wreck site between Redscar Head and Lea Lea (Lagava Iduka area) (approximate WGS84 location). Site needs to be systematically recorded in the field. CB29 55S x Ship wreckage west of Bavo Island (Dua Tanona) (approximate WGS84 location). Site needs to be systematically recorded in the field. x Site Must Not Be Damaged Submerged area of village. Site needs to be systematically recorded in the field. WWII Plane Crash x WWII Defence Facility 55S WWII Shipwreck Boera Lagatoi Wreck CB15 Tree Submerged area of village. Approximate WGS84 location. Site needs to be systematically recorded in the field. Spirit/Sacred/Story Place x x Early Missionary Site 55S Old Village (from Oral Traditions) Papa Metal (Includes Bullets) CB14 Glass Village (Submerged Area of village). Approximate WGS84 location. Site needs to be systematically recorded in the field. Animal Bones x Stone Artifacts 55S Shell Lea Lea Pottery CB13 UTM 55S CB9 Date of Site Recording ASM First Lagatoi Landing/ Stone Anchor Site Monash University Site Code Site Name Grid Reference (deleted to protect site location – see original report, David 2008) Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys (cont’d) PNG Museum and Art Gallery Site Code Table 4.5.8 x x Comments This is an important site of oral tradition. This is part of the ancestral village site of Konekaru, abandoned during the early colonial period. It is the shallow marine component now visible archaeologically. WWII pipe bridge through mangroves – (Vaihua River Region) (approximate WGS84 Location). This site needs to be systematically recorded in the field. x x B. David, B. Duncan, M. Leavesley 4-585 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project B-25D Mitchell Bomber Crash Site Site Must Not Be Damaged Recommend Stage 2 Archaeological Salvage WWII Plane Crash WWII Defence Facility WWII Shipwreck Lagatoi Wreck Tree Spirit/Sacred/Story Place Early Missionary Site Old Village (from Oral Traditions) Metal (Includes Bullets) Glass Animal Bones Stone Artifacts Shell Pottery Grid Reference (deleted to protect site location – see original report, David 2008) UTM Date of Site Recording Site Name Details of cultural heritage sites recorded during the Onshore and Offshore LNG Facilities component field surveys (cont’d) PNG Museum and Art Gallery Site Code Monash University Site Code Table 4.5.8 Comments 55S x x B-25D Mitchell Bomber (Serial # 41-30496) crash site on Bavo (Bava) Island (approximate WGS84 location). Site needs to be systematically recorded in the field. CB27 55S x x Aircraft wreck site on Piri Patch near Boera (approximate WGS84 location). Site needs to be systematically recorded in the field. CB25 55S x x Aircraft wreck site on Pullen Shoals (Hatoro Reef Group 2) (approximate WGS84 location). Site needs to be systematically recorded in the field. CB24 55S x x Aircraft wreck site between Redscar Head and Lea Lea (Lagava Iduka area) (approximate WGS84 location). Site needs to be systematically recorded in the field. CB26 55S x x A P-38 plane crash site off Boera Head (approximate WGS84 location). Site needs to be systematically recorded in the field. CB28 JD4 CB31 18-Jan-08 Lea Lea Fuel Dump 55S 55S x This is a colonial-period site consisting of a barb-wire fenceline. (GPS points deleted to protect site location). x B. David, B. Duncan, M. Leavesley 4-586 Probably the site of a WWII fuel dump north of Lea Lea (approximate WGS84 location). Further historical research required to determine exact nature of this site; site needs to be systematically recorded in the field. Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project B. David, B. Duncan, M. Leavesley 4-587 Social Impact Assessment 2008 Papua New Guinea Liquefield Natural Gas Project WWII Plane Wreck. One site, JDA17, appears to be the site of a plane crash. We were directed to this location by a Boera man, Kove Pame. Local oral tradition recounts that this is the site of a WWII plane crash. Scattered remains of the fuselage and other metal parts occur over an area some 30 x 30 m. However, much of the wreck has been taken away by local people for scrap metal. A sweep by the metal detector reveals a concentration of aluminium sheeting in grass and/or below the surface. Plate 4.5.29 Reverend Lawes and the First Church Memorial, Boera 4.5.10.1.3 Predictive Site Modelling for LNG152 Land The archaeological sites recorded during the January and April-May 2008 fieldwork include small sites consisting of isolated artifacts indicative of fleeting past human activity, to dense and extensive deposits of varied cultural materials representing permanent settlements (hamlets and villages, in particular site complexes CB1-JD1-JD2-JD3-JD5-ML3 [representing archaeologically part of Konekaru village]; and JD8-JD9-JD10-JD11-JD12-JD13-JD14 [representing archaeologically a previously undocumented ancient village] with smaller neighbouring archaeological sites undoubtedly also being closely related to these settlements). While these sites can each be treated as isolated locales of past cultural activity, many site complexes also represent functionally linked networks of sites (see Section 4.5.12.6.3 for discussion of site complexes within and near the Study Area and assessments of their significance). The difficulty is to determine which sites are contemporaneous, and how, through the course of history, they have accumulated to give the present-day archaeological record. Such an endeavour is not possible without systematic archaeological excavation and analysis, and will not therefore be attempted here. A clear pattern that has emerged from the distribution of sites is the presence of dense archaeological deposits near the present-day coastline, with numerous but significantly smaller sites further inland. Figure 4.5.29 and Table 4.5.9 show the distribution of sites by size. As Table 4.5.9 demonstrates, the blue-coded sites on the table are the sites that occur within 1 km of the coast. The green sites are those that lie between 1-1.5 km from the coast; and the brown sites are 2 those that lie more than 1.5 km from the coast. The largest archaeological sites (>1000 m ) occur on the right-hand side of the table, and the smallest on the left. There is a clear pattern of blue (close to the coast) sites on the right (large sites), and brown (inland) sites on the left (small sites). 2 Statistically, 69% (11 out of 16) of the largest sites (>1000 m ) occur within 1.0 km of the coast; all 2 of the sites greater than >1000 m in area occur within 1.5 km of the coast (i.e., none of these largest sites occur further from the coast) (Table 4.5.10). Putting the two largest site categories B. David, B. Duncan, M. Leavesley 4-588 Social Impact Assessment 2008 Papua New Guinea Liquefield Natural Gas Project together, more than three-quarters (77%), or 27 of the 35 largest sites, are found within 1.0 km of the coast, but only one (3 %) occurs more than 1.5 km of the coast. However, the inverse pattern is 2 apparent for the very small sites consisting of single artifacts, or sites less than 25 m in size. Thus only 11 % of the sites with single artifacts (e.g., a shell, pottery sherd or stone artifact) are found within 1.0 km of the coast, while 67 % of these are found more than 1.5 km from the coast. Plate 4.5.30 Missionary Tree, Lea Lea village B. David, B. Duncan, M. Leavesley 4-589 Social Impact Assessment 2008 Papua New Guinea Liquefield Natural Gas Project Table 4.5.9 Site # Size of indigenous archaeological sites containing pottery-shell-stone artifactbone remains. Site Size <1 m 2 1-25 m 2 26-100 m 2 101-1000 m 2 >1000 m JD5 x JD1 x JD2 x JD7 x JD8 x JD10 x JD13 x JD14 x JD16 x ML5 x ML10 x AAHM x AAHN x AAIP x AAIQ x ML7 x ML21 x JD6 x JD11 x JD12 x JD17 x AAIN x AAIO x ML1 x ML2 x ML6 x ML8 x ML12 x ML15 x ML18a x ML19 x ML20 x LNG3 x ML9 x AAIT JD15 x x B. David, B. Duncan, M. Leavesley 4-590 2 Social Impact Assessment 2008 Papua New Guinea Liquefield Natural Gas Project Table 4.5.9 Site # Size of indigenous archaeological sites containing pottery-shell-stone artifactbone remains. (cont’d) Site Size <1 m 2 1-25 m 2 26-100 m AAHO x ML13 x ML14 x ML16 x ML17 x ML18b x LNG2 x AAHL x AAHP x AAHR x AAHS 2 x JD3 x AAIJ x AAIM x ML3 x ML11 x AAIG x LNG1 x LNG4 x ML4 x AAHV x AAHX x AAID x AAII x AAIU x JD9 x AAIR x AAHY x AAIB x AAIS x AAHQ x AAHT x AAHU x AAHW x AAHZ x AAIC x B. David, B. Duncan, M. Leavesley 4-591 101-1000 m 2 >1000 m 2 Social Impact Assessment 2008 Papua New Guinea Liquefield Natural Gas Project Table 4.5.9 Size of indigenous archaeological sites containing pottery-shell-stone artifactbone remains. (cont’d) Site Size Site # <1 m AAIE x AAIF x AAIH x AAIK x AAIL x AAIV x 2 1-25 m 2 26-100 m 2 101-1000 m 2 >1000 m 2 Blue = 0-1.0 km from present-day coastline; Green = 1.0-1.5 km from coastline; Brown = >1.5 km from coastline. World War II and other recent sites with European-contact materials (such as the plane wreck site JDA17 and the historical fenceline site JD4) are not included. Table 4.5.10 Number and percentage of sites, by site size Distance from the coast 2 Site (m ) 0-1.0 km (‘blue’ sites) 1.0-1.5 km (‘green’ sites) >1.5 km (‘brown’ sites) # % # % # % <1 2 11 3 17 12 67 1-25 5 36 4 29 5 36 26-100 7 58 1 8 4 33 101-1000 16 84 2 11 1 5 >1000 11 69 5 31 0 0 Looking the data in another way, in total 41 cultural heritage sites were found within 1.0 km of the coast, and 22 sites more than 1.5 km away (the other 15 being found in-between). Two-thirds 2 (27 sites, or 66 %) of the sites near the coast are larger than 100 m . In contrast, of the 22 sites found more than 1.5 km from the coast, only one is of this size, while 55 % consist of single 2 artifacts only and another 23 % are less than 25 m . The pattern is very clear: large sites occur near the present coastline, while further inland (more than 1.5 km from the coast), the vast majority of sites are very small. This is consistent with a predominant coastal settlement pattern, in particular stilt villages over the shallow marine environment or on the beach-front. These large coastal sites largely correspond with the presence of sandy substrates. B. David, B. Duncan, M. Leavesley 4-592 Social Impact Assessment 2008 Papua New Guinea Liquefield Natural Gas Project 90 80 70 60 50 40 30 20 10 0 <1 1-25 26-100 101-1000 >1000 < 1- 2 6- 1 0 1- >1 0 0 1 25 100 100 0 0 < 1- 2 6- 1 0 1- >1 0 0 1 25 100 100 0 0 3 5 3 0 2 5 2 0 1 5 1 0 5 0 7 0 6 0 5 0 4 0 3 0 2 0 1 0 0 Figure 4.5.28 Percentage of indigenous archaeological sites by size. Top (blue) = 0-1.0 km from present-day coastline; Middle (green) = 1.0-1.5 km from coastline; Bottom 2 (brown) = >1.5 km from coastline. X-axis is site size in m ; Y-axis is % of archaeological sites B. David, B. Duncan, M. Leavesley 4-593 Social Impact Assessment 2008 Papua New Guinea Liquefield Natural Gas Project Figure 4.5.29 Known cultural heritage sites within and near the site security fence area, relative to distance from the coast. B. David, B. Duncan, M. Leavesley 4-594 Social Impact Assessment 2008 Papua New Guinea Liquefield Natural Gas Project B. David, B. Duncan, M. Leavesley 4-595 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project WWII Defence Sites. Two WWII defence sites have been recorded during our field surveys: CB30 and CB31. CB30 is the Boera battery previously discussed from the literature in Section 7.3.4. Douglas (2007) has conducted a previous inspection of the Boera battery, which consisted of a photographic recording of the site. An archaeological survey or site plan has yet to be completed for this site. The direction range-finding station, gun pits, magazines, and several ammunition cupboards are still intact at the site, as are the rear Section s of the cast iron gun carriages. Plate 4.5.31 Boera battery gun pit #2 Plate 4.5.32 Boera battery from north. Note range finding station (raised) and ammunition cupboards (left) and raised mound of #2 gun pit on right B. David, B. Duncan, M. Leavesley 4-596 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.33 Partially demolished magazine or accommodation quarters, Boera battery Plate 4.5.34 Electrical generator shed remains, Boera battery B. David, B. Duncan, M. Leavesley 4-597 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.35 Observation post bunker, Boera battery Plate 4.5.36 Boera battery magazine (view from south) The Boera battery is still remarkably intact, although its condition varies from when Douglas inspected the site in 2003. Two reinforced concrete gun pits with panama mounts sit atop the two highest hill crests and command an extensive view of the surrounding bay and countryside. An open range-finding tower with intact pedestal (for mounting an alidade) stands directly behind the #2 gun emplacement. Several intact concrete ammunition cupboards are located beside both gun pits and leading down the northern slope to the (ammunition) magazine which is located on an B. David, B. Duncan, M. Leavesley 4-598 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project adjacent lower ridge. The remains of a concrete pad for a generator shed are located above the magazine to the east, with another magazine or accommodation block recessed into the eastern side of this ridge. A small concrete-lined well or observation post has been dug into the northern aspect of the slope. Other structural features overgrown with vegetation were observed on two smaller hills on the seaward side of the battery (see Plates 4.5.31 to 4.5.36). It is clear from the positioning of the gun pits that they were placed to defend the northern and western entrances to Caution Bay and to repel any land attack. The 155 mm field guns mounted here were capable of launching artillery rounds up to 14 km away, and as outlined previously in Section 4.5.9.4, during regular artillery practice up to 40 rounds were fired at a time. An ack-ack (anti-aircraft gun) emplacement located at the rear of the battery (eastern side) also engaged in target practice. This would suggest that there is a strong possibility that expended and unexploded ordnances may be found within a radius of 14 km around the battery. Given that Boera village is located within 500 m of the south of the battery, it is unlikely that practice took place in that direction. Large scatters of slipped, incised and plain pottery, silcrete stone artifacts and shell middens were also encountered in the area of the northern slope of Boera Hill between the battery and the magazine, suggesting that the locality was possibly previously used for pottery production and habitation (see Plate 4.5.37). These sites were not recorded by us due to time constraints and the fact that this area lay well beyond our brief. Some of these sites are likely to have already been recorded by the PNG National Museum and Art Gallery. Plate 4.5.37 Left: Pottery rims found near Boera battery Magazine. Right: Pottery, stone artifact and shell midden scatter, Boera battery The second WWII defence site recorded during the field surveys is the fuel dump north of Lea Lea village. A telephone cable and ‘fuel hideout’ (fuel dump) were reported to the northwest of Lea Lea by Daure Veri (personal communication 2008), who also stated that ‘there is a concrete base still there, and there are posts that lead through the bush to the concrete base’. This is probably an allied fuel supply dump associated with the telegraph line shown on the 1944 military engineer’s 1 map (Lea Lea Inlet map 3968 ). This site was recorded as CB31 (see Table 4.5.8). 1 Leal Lea Inlet, 3968: 1944, Lea Lea Inlet, New Guinea. 1 inch Series, Second Edition. Netherlands East Indies Grid, Southern New Guinea Zone. NGF/010/3968 compiled by 2/1 Australian Army Topographic Survey Company from aerial photos and field observations. Reproduced by 2/1 Aust Army Topographic Survey Company. Officers using this map are requested to forward any necessary additions or amendments on the map itself and forward to A.D. Survey H.Q.N.G.F. B. David, B. Duncan, M. Leavesley 4-599 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.10.2 Offshore LNG152 Facilities Component In this section the cultural heritage sites recorded from the shallow marine environment during a review of the literature (Section 4.5.10.1) and during fieldwork (Section 4.5.10.1.2). The sites identified from the literature are listed in Table 4.5.6, and those from fieldwork in Table 4.5.8. 4.5.10.2.1 Literature Survey and Background Research Stilt Villages. No stilt village site has previously been reported from the Study Area’s shallow marine environment itself, although the ethnographically documented (presumably stilt) beach-side village of Konekaru has already been identified in the LNG152 Land survey results presented in Section 4.5.10.1 and Section 4.5.10.2 above. However, David et al. (2007:144-45) have documented the archaeological remains of stilted/raised longhouses in the inter-tidal zone in the Gulf Province to the west of the Study Area, confirming the archaeological preservability and thus possible presence of wooden settlement structures in southern PNG shallow marine environments. The extensive state of preservation of the Gulf Province sites, especially in the highly dynamic littoral zone, provides strong evidence that wooden settlement sites might exist underwater where they would be better preserved. Given what is known of ethnographic Koita and Motu cultural practices in and near the Study Area, there is a reasonable probability that fieldwork may discover the remains of submerged village sites or their associated relics, or canoe wrecks. In order for the maritime archaeological team to be able to identify such types of remains underwater, a brief visual inspection of present-day Lea Lea, Papa and Boera villages was undertaken to enable the archaeological characterisation of village constructions and their associated items of material culture. A limited photographic record was thus made of potential site components and structures, with the above aim in mind of being able to recognize archaeologically the remnants of wooden village structures (Plates 4.5.38 to 4.5.48). Particular attention was given to stilt houses and other structures such as pig pens, bridges and boat racks, for submerged wooden remains would most likely relate to such structures. Villagers accompanied the survey crew during these visits, explaining the processes of daily activities which took place at each site. This enabled the application of an ethno-archaeological approach by which to identify any new sites types that might later be discovered underwater during the surveys. Many villages around Port Moresby are still built on stilts over the sea or shallow marine environment (e.g., Boera, Hanuabada, Porebada), and although the upper superstructure of the housing is of more modern construction, based on comparisons with a limited number of historical photographs the substructure of the stilts appears in many cases to have embraced technologies also used in ancestral stilt village constructions. Shipwrecks/Navigational Beacons/Anchorages. No shipwreck surveys had been undertaken in the Study Area prior to this report. Descriptions contained in local sailing directions (Darby, 1945:234-35) discouraged shipping through the western entrance via Liljeblad passage to Port Moresby, and intensive surveys were not undertaken of this area until at least 1945. It has been noted, however, that the area was used by local coastal trader traffic to access an unmarked channel into Caution Bay that led to Port Moresby (Kidd and Neil, 1998:339). It therefore appears that any shipping through Caution Bay was probably restricted to small local traders who had intimate knowledge of the area. Although it is unlikely that any major shipwrecks would have occurred in this region, as it was not frequented by passing traffic, the possibility exists that a vessel may have tried to take shelter in the region in adverse weather (and was subsequently wrecked). No navigational charts located showed any evidence of in-water navigational beacons or anchorages for this area, which is not surprising given the shoal nature of the Bay. No references to recommended anchorages or sailing routes were found for the Caution Bay area. By implication, it is therefore unlikely that navigational beacons had been installed in the area. The possibility does B. David, B. Duncan, M. Leavesley 4-600 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project exist, however, that isolated anchors (associated with vessels seeking shelter in adverse conditions) may have been lost in this area. Plate 4.5.38 Boera stilt village, February 2008 Plate 4.5.39 Traditional stilt house, Lea Lea village south. Note canoe platform at front B. David, B. Duncan, M. Leavesley 4-601 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.40 Close up of the underside of a stilt house at Lea Lea village Plate 4.5.41 Canoe skids, Papa village B. David, B. Duncan, M. Leavesley 4-602 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.42 Net drying rack, Boera village Plate 4.5.43 Pig pen on stilts, Lea Lea village Plate 4.5.44 Chicken coup under stilt house, Lea Lea village south B. David, B. Duncan, M. Leavesley 4-603 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.45 Outrigger canoe, Papa village Plate 4.5.46 Canoe rack, Papa village Plate 4.5.47 Fisherman repairing nets, Lea Lea village Plate 4.5.48 Dugout canoe, Lea Lea village south Plane Wrecks. Based on the literature review presented in Section 4.5.10.1.1, 10 plane wreck sites are known from vicinity of the Study Area, although none have been recorded to occur within the Study Area itself. Seven of these sites crashed in the sea or mangroves, or are missing and believed to have crashed in or close to the sea. They include a B-24 Consolidated Liberator which crashed in the mangroves near Caution Bay; a B-25 Mitchell Bomber which crashed 1 mile west of Bava Island; a B-25c Mitchell Bomber crashed at Hidiha reef; an A-20 Havoc Bomber, believed to have crashed in the sea between Bavo and Hidiha Islands; a P-39 Airacobra, which is thought to have crashed in the mudflats 25 miles northwest of Port Moresby; a P-39 F-1 Airocobra last seen at Redscar Bay; and a B-25 H-1 Mitchell Bomber which crashed in the sea near Redscar Head. A number of these plane crashes involved the death of pilots or other personnel (e.g., the B-24 Consolidated Liberator; the A-20 Havoc Bomber; the P-39 F-1 Airacobra; the B-25 H-1 Mitchell Bomber); they may therefore be war graves. These sites have been listed in Table 4.5.6. Telegraph Cable. Site CB5 consists of a telegraph cable line. It was not inspected, as it was only identified from historical records after fieldwork had been undertaken. Further research is required to identify its exact location, and to ensure the facility is not still being used. It is included in Table 4.5.6. B. David, B. Duncan, M. Leavesley 4-604 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.10.2.2 Field Survey Side-Scan Sonar. Several anomalies were recorded during the side-scan sonar survey, but none resembled expected signatures for aircraft or shipwrecks. In all cases where the water was less than 5 m deep, the anomalies proved to be coral reefs. An example of a typical coral outcrop detected by side-scan sonar is shown in Plate 4.5.49. Plate 4.5.49 Example of side-scan sonar image of a coral heads (‘bommies’). Note the large shadow behind the coral patch. The side-scan recorded a predominantly flat and featureless bottom for the deeper seabed areas (5-15 m). The potential ‘sites’ identified in deeper water were investigated using a drop camera (due to problems with constantly deteriorating weather during planned diving days). Although no cultural material was located using this method, it is possible that underwater surveys using divers would have yielded more detailed inspection results for those locations. A table of detected underwater anomalies is presented below (Table 4.5.11). As noted, none of the anomalies detected here resembled the anticipated readings expected for ship or plane wrecks. Location ID10 was not inspected as it occurred in Option Area 2 and adverse weather restricted access. This location should be inspected by divers prior to developments to determine if it is archaeological in nature (see recommendation in Section 4.5.14). B. David, B. Duncan, M. Leavesley 4-605 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.11 Possible side-scan sonar anomalies identified during the survey. ID site # 1 2 Run 1 1 Easting (has been deleted to protect site) Northing (has been deleted to protect site location) Date (March 2008) Comments Observations 9 th Possible coral bommie coral reef 9 th Possible coral bommie – high profile coral reef Possible coral bommie – high profile extent with 3b coral reef 3 3 9 th 4 3 9 th Possible coral bommie – long thin pile or canoe float? coral reef 5 3 9 th Faint trace – possible buried mound? nothing located 9 th Possibble coral bommie – check coral reef 6 7 8 3 1 2 10 th Possible hit 10 th Small solid hit not located Small solid hit not located 9 3 10 th 10 4 10 th Large shadow not checked – in Option 2 area 11 10 10 th Small hit nothing located 10 th Long straight hit nothing located Small shadow coral High shadow coral 12 10 13 5 12 th 14 7 12 th B. David, B. Duncan, M. Leavesley 4-606 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project B. David, B. Duncan, M. Leavesley 4-607 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Magnetometer: Effects of Local Magnetic Anomalies on Maritime Remote Sensing Survey. Upon commencement of the survey, it was discovered that there was an extremely high localised background magnetic clutter reading across most of the Study Area. This was probably associated with either an igneous rock substrate under the seabed; materials derived from the breakdown of fine-grained (but highly magnetic) volcanic/igneous rock (e.g., basalt) transported to the region by rivers as erosion products (in the form of fine silt or clays); or materials from volcanic ash/particles falling through water. These sediments may subsequently form a depositional lens of highly thermo-remnant material on the seabed in the form of fine silt or clay layers (Gibson and George, 2003: 91-97). All of these materials can produce high magnetic readings. Despite adjustment of the equipment to try to tune out these anomalies, the background readings were so high that they were effectively peaking out the sensitivity range needed to detect archaeological sites. The magnetometer survey of this area was severely hampered by high localised magnetic background readings which would effectively mask the readings of smaller magnetic anomalies (including cultural materials such as bombs or other smaller items such as concentrations of ceramic sherds). Similar circumstances have been experienced by Brad Duncan during other magnetometer surveys at archaeological sites around Australia. Despite the background magnetic clutter, we persisted with the magnetometer survey for the entire period under the premise that a larger object may give a readable signal. As a result of these geological interference and magnetic susceptibility, no small archaeological sites could be identified in the area using the magnetometer. However, a test of the magnetometer equipment to detect a large iron shipwreck (the SS Mac Dui) in Port Moresby harbour showed that the equipment would still detect large metallic anomalies. It is therefore likely that no large anomalies such as shipwrecks exist in the surveyed area. No sites were identified using remote sensing equipment inside the Study Area. This was consistent with oral and social knowledge of this area collected from local fishermen and divers, who did not indicate any knowledge of any submerged sites in this area. However, it is possible that sites may be buried below the seabed, but are currently not visible. Traditional Story Sites. Four traditional sites associated with the story of the ancestral hero Edai Siabo and his first lagatoi (see Section 4.5.7.9 for details) were located in the Caution Bay area. Two of these sites (CB7 and CB8) have already been presented under the LNG152 Land component in Section 4.5.10.1.2 above; the other two sites (CB6 and ASM) occur in the shallow marine environment and are therefore presented here. Each of these four sites is an integral component of the first lagatoi story. They were each inspected during the course of our fieldwork, even though they lay outside the Study Area. Each of these sites is of the utmost cultural significance, relating to what is arguably the most important customary oral tradition of the Western Motu. Moi and Mea Dobi (personal communication 2008) related the following story: This beach is associated with the [story of the] first lagatoi canoe. That anchor is where Edai Siabo from Boera first came ashore. There are underwater caves at Hidiha [Idihi] Island. He was pulled into an underwater cave by sea spirits (or ancestors) and they taught him how to build the first lagatoi canoe. His mates saw his legs sticking out from the sea, and pulled him out of the cave. He later made a model of a lagatoi, but his mates laughed at him. He then made a full scale model of it, which was the first large lagatoi canoe. They were hard times then, so he went to Kerema and established the hiri trade. He built the first lagatoi on the beach at Apau, which was the village before Boera. He sailed in around to here [First Lagatoi Landing/Stone Anchor site ASM], and threw in the anchor here. The anchor was left where he came ashore. This is the location of the sacred stone anchor from the first lagatoi boat [Moi pointed to a round, light grey circular stone approximately 60 cm thick and 45 cm in diameter). B. David, B. Duncan, M. Leavesley 4-608 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project This is a traditional place for us, and we do not disturb the anchor. One time a researcher (name not recorded) came and tried to take a piece of the anchor, you know to see what rock type it was, but the bees came and stung him and scared him off. Site CB6 is the site of the sea cave in which Edai Siabo was instructed about the making of the first lagatoi by the spirit-being. It lies approximately 50 m offshore to the southeast of Hidiha Island. Moi Dobi (Boera fisherman) pointed out its location. The cave mouth is set in a shallow reeftop in water less than 1 m deep. No features of the cave could be discerned during an inspection of the site, due to it being currently silted up with sand. ASM was traditionally the location where the first lagatoi came ashore and threw over its stone anchor. The basalt anchor remained in this location, and is still visible at low tide (see Plates 4.5.50 to 4.5.52). The anchor is roughly circular in shape, approximately 60 x 45 cm in thickness and diameter. The anchor is possibly of a type designed to fit in a cane basket, which was then attached via ropes to the vessel. Similar stone anchors were observed by missionaries in 1883 and were often attached to boats by 100 fathoms of line (e.g., Lennox, 1903:1). One of the archaeologists on our team (Lyall Mills, personal communication 2008) has observed similar designs in use on canoes in Fiji. The beach in this area has high concentrations of ceramic sherds scattered over a very large area. High concentrations of brown silcrete stone artifacts (cores and flakes) along this beach were also identified by Mea Dobi (personal communication 2008) as ‘Kavari’ which were used to make shell bands, a practice which ended in the 1960s. Plate 4.5.50 Location of First Lagatoi Landing/Stone Anchor site, Boera (marked by arrow) B. David, B. Duncan, M. Leavesley 4-609 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.51 Stone artifact on the beach of the First Lagatoi Landing/Stone Anchor site. ‘Kavari’ artifact purportedly used to make shell armbands Plate 4.5.52 First Lagatoi Landing/Stone Anchor site, stone anchor at low tide near Boera village (exposed stone = 40 x 60 cm) Lagatoi Wrecks. There were three reports of possible lagatoi wrecks in Caution Bay. Lea Lea Councillor Hene Totana (personal communication 2008) indicated that he knew the location of a lagatoi stone anchor. It was two miles north of Lea Lea village. The lagatoi was undertaking a voyage from Lea Lea to Port Moresby when it sank. The anchor from the wreck was found and dragged ashore onto the beach. The locals reported that there was nothing left of the wreck as it was likely that the hull had floated off. The lagatoi was carrying sago, mats and possibly pots for barter, although the latter were probably smashed up during the wreck. He could not arrange to take us to the site as the bad weather prevented access, but was happy to arrange this at a later date. Auda Delena (personal communication 2008) has marked the location of the stone anchor at Gabi to the north of Lea Lea on the traditional names map. This site was recorded as CB16 (see Table 4.5.8 for details). B. David, B. Duncan, M. Leavesley 4-610 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Gau Ario (personal communication 2008) noted that local fishermen had reported the possible finding of a lagatoi wreck site to the southwest of Papa. Although further information about this site was being sought by Ario, it was not available when this report was being finalised. This site was recorded as CB17 in Table 4.5.8. Gau Ario (personal communication 2008) also reported that fishermen knew the location of another lagatoi which sank on a voyage from Koiruru to Kido Point in either the late 1800s or early 1900s with a cargo of sago and pots. It sank and was pulled ashore at Kerama. This site was recorded as CB18 (Table 4.5.8). Given that Boera was a major centre of hiri trade, it is possible that other lagatoi wrecks may exist in the Caution Bay area. As it is unlikely that the timbers from these vessels would have survived intact above the seabed (due to the presence of teredo worms and other marine borers), the only probable remains of such sites above the seabed would be as pots or ceramic sherds. Given the high background magnetic readings for this area, it may be unlikely that their locations can be discovered using a magnetometer. Such sites would be of very high cultural significance if ever discovered. Former Over-Water Stilt Village Sites. Villagers from Lea Lea, Papa and Boera each reported the loss of coastline in modern times. This would suggest that the coastline in this area is highly mobile and subject to long-shore drift. This has a number of implications for potential archaeological sites across this region. It is highly probable that former village Section s may now exist underwater and are possibly covered with sand/sediment/mud on the seabed. The high mobility of sediments in this area may also cause other archaeological sites to be periodically covered and uncovered. Such sites would not be evident when using the remote sensing techniques employed in this study if they were covered at the time of the survey and if they did not contain very high concentrations (rather than just spread) of ceramics detectable through magnetometer surveys. Furthermore, it is probable that other sites along the foreshore may now be buried under prograding shorelines where sediments scoured from other areas have now been deposited. Nevertheless, it is apparent from visiting current villages that older Section s of these villages evidence archaeological deposits and/or oral traditions which testify to their prior location (see Table 4.5.8). Site CB13 is a submerged, past Section of Lea Lea village. Lea Lea villagers reported that the former location of their village was up to 200 m further to sea than the present shoreline, and that several houses had collapsed or been washed away due to coastal change in recent years. Auda Delena (personal communication 2008) indicated that this former village site was located on the northern bank of Lea Lea Creek to the west of the current village. CB14 is an old, historical part of Papa village. At Papa, the villagers reported similar effects on former stilt villages located over water. A photograph from WWII shows that the southern parts of the village were formerly mounted on stilts over the water (AWM Photos #060903 and 060904 – see Plate 4.5.53). As was the case at Lea Lea and Papa, at Boera residents also reported that the coastline is being washed away and that those houses on low ground are being threatened with inundation. The inter-tidal areas to seaward of the current stilt houses built over the sea at Boera were inspected at low spring tides. An extremely dense deposit of compacted pottery sherds was evident along the entire Boera village coastline up to 50 m beyond the current housing; we have recorded this archaeological site as CB15 (see Plate 4.5.54). This would suggest that these remains may be of an earlier phase of the village which extended further out across the water, as pottery manufacture has largely (but not entirely) ceased at Boera. It should also be reiterated that Boera was ethnographically a renowned centre of pottery-making for hiri (as well as local and regional) trade, and the site could therefore be of some antiquity. B. David, B. Duncan, M. Leavesley 4-611 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Closer to the Priority Area, a search of the offshore areas immediately adjacent to Konekaru beach did not reveal any indications of a former pier or jetty structure. Although the author hypothesised that any former pier may have been built of galvanised pipes (as suggested by the initial discovery of a single iron pipe in the inter-tidal zone during the terrestrial survey), a later closer inspection of the underwater seabed discounted this possibility as no other similar relics were located and there were no indications of any depressions in the coralline seabed which might be caused by such structures. However, a marked increase in coastal erosion noted by many clans in the area may explain this, as the area may have formerly been covered with sand/silt sufficient to bed a structure but which is now removed. A number of WWII bullet casings (22, 303 and 50 mm calibres) and other domestic equipment (spoons and batteries) were also discovered on depressions in the coral seabed approximately 2-5 m from the high tide mark, indicating the presence of a colonial-period site here (see Plate 4.5.55). Plate 4.5.53 Stilt houses at Papa village, 1943. Left: AWM photo 060903. Right: AWM photo 060904 Plate 4.5.54 Dense pottery scatter over the entire intertidal zone to seaward of Boera village at extremely low tide (note that the site extent covers all areas seen in this photograph). Dense scatter of pottery sherds evident in foreground B. David, B. Duncan, M. Leavesley 4-612 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.55 Historic artifacts discovered underwater at Konekaru beach (303 bullet casings and spoon) Despite the absence of structures, the existence of this beach amongst such dense stands of mangrove, along with the presence of new mangrove shoots (even on coral substrate), suggests that this area was deliberately cleared and maintained, possibly as a beach area. The presence of the dense clusters of pottery, shell middens and other historic artifacts onshore associated with sites JD1, JD2 and JD3 (directly to the east of the beach, which were discovered during the terrestrial survey; see Section 4.5.10.1.2 above) suggest that the area is probably associated with a historic former village site at this location, and this is confirmed by oral traditions of the presence of Konekaru village in this area. Stuart (1973:277) has previously documented similar cleared and maintained areas used to launch/land canoes in other nearby areas closer to Port Moresby. The archaeological materials in the shallow marine environment in this area have been recorded as site CB1. A search of this area at low spring tides when the area was dry revealed a light concentration of unslipped plain pottery (approximately 50 pieces; many pot rims) scattered on a hard coral and sand substrate around the edges of the mangroves lining Konekaru beach, particularly on the northern side (Plate 4.5.56). Stone artifacts are also present (e.g., Plate 4.5.57). This scatter was concentrated predominantly on the northern seaward edge of the mangroves lining the beach up to 150 m offshore, and as far as 80 m to the north and 150 m to the south, where the substrate turned to mud. It is possible that other artifacts lie buried in the softer seabed further along from these locations. Closer inspection of the interior edges of the mangrove swamps surrounding Konekaru Beach also revealed a number of historic period bottles which have been th th dated to around the late 19 or early 20 century (see Plate 4.5.58). B. David, B. Duncan, M. Leavesley 4-613 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.56 Pottery rim from the north side of the inter-tidal zone at Konekaru beach Plate 4.5.57 Stone core artifact from the inter-tidal zone at Konekaru beach th Plate 4.5.58 20 century medicine bottle found among the mangroves on from the north side of the inter-tidal zone at Konekaru beach Along the beach front approximately 3 m from the high tide mark, the sandy shore gave way to a substrate of compacted coral, in which was embedded a single pottery sherd. Although it could not B. David, B. Duncan, M. Leavesley 4-614 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project be positively determined if the coral had grown over the sherd, or if the sherd had been simply wedged into the coral, the former possibility would suggest that the site was of some antiquity. Gau Ario (personal communication 2008), a Papa resident and member of the LNG152 Land survey team, recalled that the area had once been used by paratroopers during the war. This observation is consistent with the discovery of several 303 rifle and 50 mm bullet shells in shallow water on the edge of Konekaru beach. The concentration of artifacts both on and offshore may suggest that either a stilt village site which extended over the water was formerly situated in this locality, or that the site was used in association with shipping produce from the sisal plantation documented in Section 4.5.9.3. In the case of the latter scenario, given that there appears to be no evidence of a pier at this site, produce may have been loaded in canoes for transport to market or lighterage to larger vessels offshore. However, the site is in the area of the ethnographically-documented Konekaru village, and geographically adjacent to archaeological site JD1+2 which has been documented to represent part of the archaeological remains of this village (e.g., Section 4.5.10.1.2). This shallow marine site is almost certainly also part of Konekaru village. WWII Sites: Pipe Bridge through Mangroves. The remains of an allied army pipe bridge were reported in the mangroves between Boera and Papa near the Vaihua River (site CB3). Mea Dobi (personal communication 2008) indicated that this facility was used to transport munitions through the mangroves, and the remains had been seen by women who were collecting mud crabs. This site was not inspected due to landward access restrictions imposed by the discovery of unexploded ordnances in the LNG152 area (see Table 4.5.8 for details). Plane and Ship Wreck Sites. Two plane crash sites were recorded from the shallow marine environment during our field surveys. Site CB21 is a Japanese Zero crash site. A number of fishermen at Lea Lea village told us that a single engine Japanese aircraft had crashed on the mud bank at the southern extremity of Lea Lea village (south of Lea Lea Inlet). They maintained that the pilot escaped the wreckage and was held by local villagers until the Allied Officer arrived to collect him (Gaudi Tua and Joe Delena, personal communication 2008). A family living close to the wreck site held parts of a wing strut and instrument panel from this plane wreck, collected when the plane crashed on the southern bar entrance to Lea Lea village (see Plate 4.5.59). They explained that after the plane had crashed it had sunk ‘underground’ into the mud, although the engine and wing were supposedly still visible at the site. The family then led Brad Duncan offshore at low tide onto the black sands of the bar. Although they indicated the approximate position of the plane wreck (which was approximately 100 m offshore in a direct line from their house), the falling tide meant the wreckage was under breaking waves and could not be reached. Brad Duncan was assured that the remains were still visible on the falling tide and he could be guided exactly to them by other fishermen if he returned later (Robert and Tauri Auri, personal communication 2008). The site is marked on the traditional names map at a place called Kirima Iduka (Auda Delena, personal communication 2008). The site is probably the Japanese Zero (A6M2 Model 21 Zero – Manufacturer # 3537) which is described historically (see Section 4.5.10.1.1 above; Table 4.5.6) as having crashed at Boera. The location of this site has yet to be verified, but there seems little doubt that an aircraft wreck does exist in this immediate vicinity. B. David, B. Duncan, M. Leavesley 4-615 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Plate 4.5.59 Instrument panel and aluminium skin from Japanese Zero fighter at Lea Lea Site CB22 was shown to the field survey team by Moi Dobi (Boera village) during the inspection of the First Lagatoi Story Sea Cave site (CB6). It is the wreck of a B-25C Mitchell Bomber (see Table 4.5.6 for details). The plane lies in less than half a metre of water at low tide and is spread over an area of at least 30 x 200 m. A 16 x 2 m Section of the main fuselage lies intact (but upside down) on the shallow reeftop approximately 300 m east of Hidiha Island. The retracted wheel struts and landing gear are still intact and in remarkably good preservation, with the chrome wheel rams still shining. This would indicate that the aluminium Section s of the plane are acting as an anode and corroding in preference to the more inert metals. Approximately 10 m to the northeast are remains of the cockpit flying controls. A 14 cylinder aluminium rotary engine lies approximately 200 m to the east and is awash at high tide. Other Section s were evident above water but could not be easily reached because of the low tide. Although the site was intact when the plane landed, it has obviously been broken apart by large oceanic swells which break across the reef at high tide. Seven other aircraft crash sites were reported during oral histories, but their actual locations have not been positively identified: • CB23: Aircraft Wreck – Between Redscar Head and Lea Lea (Lagava Iduka area) (approximate WGS84 location: GPS points deleted to protect site, see original report – David 2008). Lea Lea fishermen (Daure Veri, Nou Henau, and Morea Mapore, personal communication 2008) reported that pieces of an aeroplane wing had been pulled up in a net by fishermen close to the southeastern side of Redscar Head. Fishermen knew the location of the wreck, which was considered a good fishing spot. Nou’s brother-in-law (Heni) has a bearing for the plane, which uses the transit marks of a coconut tree and mountain, plus the colour of the reef, and has given an undertaking to help relocate the wreck (Nou Henau, personal communication 2008). Mea Dobi (personal communication 2008), a Boera fisherman, also placed the aircraft wreck in this area in the vicinity of Kohua. This site location has been marked on Figure 4.5.31 at the Lagava Iduka area. Elders of Lea Lea village recalled seeing a plane crash into the sea followed by a loud explosion as it burst into flames when it hit the water during WWII (Auda Delena, personal communication 2008). The wreck is probably the B-25 H-1 Mitchell Bomber (Serial # 43-4341) which was reported to have crashed into the sea near Redscar Head in 1944 (see Section 4.5.7.9, Table 4.5.6 above). B. David, B. Duncan, M. Leavesley 4-616 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project As airmen were possibly killed when the aircraft crashed, this site should be treated as a war grave. • CB24: Aircraft Wreck – Between Redscar Head and Lea Lea (Lagava Iduka area) (approximate WGS84 location: GPS points deleted to protect site, see original report – David 2008). Another aircraft crash site reported by a local diver as being situated in 18 m of water has been identified as a P-38 (John Miller, personal communication 2008). This may be the P-39 F-1 Airacobra (Serial # 41-7136) which was last seen over seen Redscar Bay in 1942 (see Section 4.5.10.1.1, Table 4.5.6 above). If Miller’s location for the site is correct, it will lie in the path of the proposed shipping channel. As airmen were killed when the aircraft crashed, this site should be treated as a war grave. • CB25: Aircraft Wreck – Pullen Shoals (Hatoro Reef Group 2) (approximate WGS84 location: 492025 E 8968388 N). Lea Lea fishermen have reported the remains of a metal frame and wings discovered in the middle of Hatoro Reef Group 2 on the edge of the deep side (Auda Delena, personal communication 2008). The reef is marked on Aus Chart 379 as part of the Pullen Shoals. This wreck may be the remains of the A-20G Havoc (Serial # 43-9122) which crashed between Hidiha and Bava Islands in 1944 (see Whiting, 1994:137) (see Section 4.5.10.1.1, Table 4.5.6). Further work is required to verify the location of this site. As airmen were probably killed when the aircraft crashed, this site should be treated as a likely war grave. • CB26: P-38 – Off Boera Head (approximate WGS84 location: GPS points deleted to protect site, see original report – David 2008). A P-38 aircraft wreck was reported to John Miller (personal communication 2008) in the vicinity of the southwest of Boera Head in approximately 1-4 m of water, although he has not seen the site himself. Its existence is yet to be verified. • CB27: Aircraft Wreck – Piri Patch – Near Boera (approximate WGS84 location: GPS points deleted to protect site, see original report – David 2008). Daro Avei (personal communication 2008), a Boera fisherman, has reported plane wreckage on the Piri Patch reef in about 5 m of water. This may be the same aircraft (P-38) reported by John Miller off Boera Head (site CB26 above). Its existence is yet to be verified. • CB28: B-25D Mitchell Bomber (Serial # 41-30496) – Bavo Island (Bava Island) (approximate WGS84 location: GPS points deleted to protect site, see original report – David 2008). This site was reported by a number of divers and fishermen (Moi Dobi, Neil Whiting and John Miller, personal communication 2008). The wreck lies in 3 m of water approximately one mile due west of Bavo Island (traditional name Bava island), with parts of the fuselage, cockpit and propeller intact. The propeller is visible at low tide. The site was not inspected during fieldwork. As airmen were killed when the aircraft crashed, this site should be treated as a war grave (see Section 4.5.10.1.1, Table 4.5.6). • CB29: Ship Wreckage – West of Bavo Island (Dua Tanona) (approximate WGS84 location: GPS points deleted to protect site, see original report – David 2008). Large Section s of a wrecked ship lie in deep water about three miles west of Bava Island at Dua Tanona (Moi Dobi, personal communication 2008). It is probable that this site is the remains of the B. David, B. Duncan, M. Leavesley 4-617 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project motor launch sent to rescue the crew of the B-25D Mitchell Bomber (Serial # 41-30496) which crashed near Bava Island (see Whiting, 1994:126-27) (see Section 4.5.10.1.1). Although a small recreational fishing boat was supposed to have wrecked close to the LNG152 area (Maurice Brownjohn, personal communication 2008), no further information was available to indicate the precise location of this vessel. 4.5.11 Modelling the Distribution of Cultural Heritage Sites in the Study Area and Site Perimeter Fence Area In sum, 162 cultural heritage sites were recorded in the field and from the literature within and near the Study Area. The list of all the recorded cultural heritage sites is the combination of Tables 4.5.6 and 4.5.8. The total list of sites includes 10 site types (three individual sites in that list are duplicated as they belong to more than 1 category: two clay sources that are both traditional resource sites and indigenous archaeological sites; and the 1st lagatoi anchor site, that is both a traditional story/sacred site and an indigenous archaeological site): • Indigenous archaeological sites (N = 120). • Traditional resource sites (N = 3). • Traditional story/sacred sites known from oral traditions (N = 5, excluding the traditional village sites below that are also story and/or sacred sites). • Traditional village sites known from oral traditions (N = 12). • Lagatoi wrecks (N = 3). • Early missionary sites (N = 2). • WWII defence facilities (N = 3). • WWII aircraft crash sites (N = 11). • WWII shipwrecks (N = 1). • Other colonial period infrastructure sites (N = 2, one of which contains distinctive components). Within the Site Perimeter Fence Area as defined in Section 4.5.5.1 above, to date 63 archaeological sites (and no oral tradition sites) have been recorded. This consists of both small and large sites, the latter including the archaeological village site represented by the very large SC5. These sites will be further discussed in Section 4.5.12 (Significance Assessments). What we do not yet know from these sites is the quantity and distribution of sites within the Site Perimeter Fence Area as a whole. In this Section, extrapolations are made from the data presented above to estimate the number of cultural heritage sites present within the entire site security fence area. To do so, the data obtained from the 100 % archaeological surveys undertaken in January 2008 and those from the transect surveys of April and May are used, rather than all of the known sites recorded from the literature and field surveys. The reason for using the field data only is that it is contiguous (and largely overlaps) with the site security fence area, and it represents the only data obtained systematically from ground surveys. The distribution of sites within this area relative to the coastline can thus be modeled as a way of predicting the overall distribution of sites within the site security fence area. However, it is stressed that this will only give a general indication of the types and quantities of sites likely to occur within the site security fence area. This approach cannot reveal the specific location of individual sites; nor is it likely to reveal the occurrence of rare site types that remain unrepresented in the Study Area upon which the modelling is founded. Furthermore, we refrain from breaking-down the estimate of overall site numbers into smaller constituent components (e.g., B. David, B. Duncan, M. Leavesley 4-618 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project number of sites on sandy sediments; number of sites at given distances from the coast or from creeks) because of the generally small area at stake (especially for inland Section s where ground surveys were largely based on 3 m-wide transects), whereby extrapolated calculations would exaggerate non-representative sample areas (this is often a significant limitation in surveys that do not survey all component parts in comparable degrees of intensity – in this case a product of limited surveys covering the inland areas as a result of restricted access due to the discovery of UXOs). Complete (100 %) field surveys will thus be needed within the entire area of the site security fence prior to commencement of developments (see Section 4.5.14 recommendations). 2 During the January and April-May field surveys, a total of 1.70 km of the Onshore LNG Facilities component was systematically surveyed (equal to 21.4 % of the area within the site security fence). Within this area, 78 archaeological sites were found (plus the oral tradition site of Konekaru, which is represented archaeologically in the 78 archaeological sites and will not therefore be further considered in the following extrapolations). These include 20 small sites consisting of isolated or small sets of four or less unstratified cultural items lying on the ground surface; 11 small stratified 2 2 sites 25 m or less in area; 12 medium-sized sites between 26 and 100 m in area; 19 large sites 2 2 between 101 and 1000 m in area; and 16 very large sites over 1000 m . As shown in Section 4.5.10.1.3 above, archaeological sites in this area can be divided into three distinctive geographical groups: those within 1.0 km of the coastline (which tend to be most numerous and the largest); those between 1.0 and 1.5 km of the coast; and those more than 1.5 km from the coast, which tend to be more sparsely distributed, less dense in cultural items and small in size. While all the very large sites occur within 1.5 km from the coast (and most – 77 % – large and very large sites recorded come from within 1.0 km of the coast), inland areas contain a very large number of very small sites consisting of a handful of artifacts (often only one or two); the larger of the inland sites also always contain very sparse distributions of cultural materials (typically 2 <1-3 m ) whereas closer to the coast sites usually have very high densities of artifacts (often 2 measured in the hundreds or more per m ). This is taken to indicate that people often used the hinterland for travel, gardening, hunting, short-term camps and the like, but within the LNG Facilities site area generally settlements were almost exclusively established along the coast (consistent with the ethnographic evidence; see Section 4.5.7). B. David, B. Duncan, M. Leavesley 4-619 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.30 Known cultural heritage sites within and near the site security fence area B. David, B. Duncan, M. Leavesley 4-620 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project B. David, B. Duncan, M. Leavesley 4-621 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 2 Table 4.5.12 shows the areas (in km ) covered for each of the three zones during the January and April-May field surveys. Table 4.5.13 shows the number of sites recorded during these field surveys, by distance from the sea and site size. We do not extrapolate from these data for the Site Perimeter Fence Area as a whole here, largely because the narrow, 3 m-wide transects upon which many of the >1.5 km areas in particular were surveyed are not well amenable to this kind sub-divisional extrapolation. This analytical limitation was a product of the unexpected discovery of UXOs which impeded our initial 100 % surveys (in January), and subsequently further unexpectedly impeded our April-May surveys when we had planned to complete the 100 % surveys but once we arrived in the field were unexpectedly told we could not as the area had not yet been cleared of UXOs (and therefore were required to survey along transect lines only). Because of this, here we simply extrapolate the total number of sites for the Site Perimeter Fence Area based on the total number of sites found within the surveyed areas. We stress the general point that the largest and densest sites found are invariably located on sandy substrate sediments (typical of coastal and river-bordering environments) rather than on hard, clayey sediments (typical of hinterland environments). Table 4.5.12 Size of areas covered during the January and April-May 2008 field surveys, by distance from the coast Distance from the coast 0-1.0 km 1.0-1.5 km >1.5 km Area surveyed (in km ) 1.37 0.31 0.02 Area surveyed (in ha) 137.3 31.0 2.0 2 Table 4.5.13 Number of archaeological sites within the January and April-May 2008 survey area, by distance from the sea and site size Small, unstratified surface sites 1-25 m 0-1.0 km from coast 2 5 7 16 11 1.0-1.5 km from coast 4 3 1 2 5 >1.5 km from coast 14 3 4 1 0 Distance from coast Stratified or Likely Stratified Sites 2 26-100 m B. David, B. Duncan, M. Leavesley 4-622 2 101-1000 m 2 >1000 m 2 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Figure 4.5.31 Known cultural heritage sites within and near the study area B. David, B. Duncan, M. Leavesley 4-623 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project B. David, B. Duncan, M. Leavesley 4-624 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 2 The area within the site security fence consists of a land area 7.93 km in size. This consists of 2.04 2 2 2 km of land within 1.0 km of the coast; 1.64 km from 1.0 to 1.5 km; and 4.25 km more than 1.5 km. 2 Based on the site recovery rate from the January and April-May survey area (78 sites from 1.70km ), this means that 364 sites can be estimated to occur within the site security fence area as a whole. Based on the survey data, a significant proportion (~40 %, or 146 of these sites) of these are extrapolated to be very small hinterland sites (single artifacts or small groups of artifacts). Most of the very large sites are likely to have been found already, as their considerable spatial extent means that they are likely to have been found across much of the 100 % area as well as the transect surveys, although complete surveys of the entire site security fence area will need to be undertaken to determine the exact location of all sites across the landscape prior to any developments proceeding (see Section 4.5.14). These figures will help inform the management recommendations for the proposed LNG Facilities developments presented in Section 4.5.14. B. David, B. Duncan, M. Leavesley 4-625 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.12 Significance Assessments Assessing the significance of cultural heritage sites is fundamental to cultural heritage site management planning (see Moratto and Kelly, 1978; Pearson and Sullivan, 1995; Smith 1996). Statements of significance can be made in respect to specific sites or places, or to a grouping of sites/places within a circumscribed area. In the case of the latter, the importance of a site complex, cultural heritage area or precinct may be greater than the sum of its individual sites. This is addressed explicitly in section 4.5.12.6.3 below. Cultural heritage significance is the value of cultural heritage sites or places to society (Kerr, 1990:3). The major criteria by which the significance of cultural heritage places is usually assessed can be divided into five types: 1. Social and Political 2. Scientific 3. Historical 4. Educational and Economic 5. Aesthetic Each of these significance criteria can be assigned a relative value from low to very high at the Local, Provincial or National level. For each assessment criterion, any given site or site complex can thus be allocated a relative significance value for each level. Such a process of significance assessment forms the basis of the Burra Charter (Australian ICOMOS charter for the conservation of places of cultural significance), which is widely employed by heritage consultants across the Pacific region (MarquisKyle and Walker, 1992). 4.5.12.1 Social and Political Significance If a place has importance for a particular cultural or ethnic group – either a majority or minority group (Lennon, 1992:4) – for religious, spiritual, or other symbolic reasons it has social significance (Johnson, 1992; Moratto and Kelly, 1978:10). Places of social significance are usually important in maintaining a community’s integrity and sense of place; that is, a sense of belonging to a particular area as a distinctive cultural group with a distinctive spiritual connection (Hall and McArthur, 1993a:8; Hodges, 1993; King et al., 1977:96). For many peoples, indigenous archaeological sites (e.g. burials) and European-indigenous contact sites (e.g. missions, plantations) have strong social significance. In recent years and in many places across the Pacific, such associations have become increasingly political as indigenous peoples regain control of their ancestral lands and re-establish senses of place following the period of European colonial rule (see Boyd and Ward, 1993:112). An example of a cultural heritage site with very high social and political significance is the Iatmul haus tambaran at Kanganamun village, East Sepik Province. This site is listed on the PNG National Museum and Art Gallery register as a National Cultural Property because of its great social, political and symbolic significance at Local, Provincial and National levels. Other PNG examples of cultural heritage sites with very high social and political significance include Kuk in the Wahgi valley (Western Highlands Province), the Kokoda Trail (Northern Province) and Samarai Island (Milne Bay Province) (these sites are also renowned for other significance criteria – see below). 4.5.12.2 Scientific Significance The scientific significance of cultural heritage places represents their ability to furnish data on, and insights into, either the history of cultural activities (social, technological and ecological) and/or the history of natural/environmental conditions (see Bickford and Sullivan, 1984; Moratto and Kelly, 1978; Pearson, 1984). For example, archaeological sites provide information on past activities, particularly B. David, B. Duncan, M. Leavesley 4-626 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project everyday lifeways, which are often not always available in documentary sources or oral traditions. Such insights apply equally to societies with and those without writing. Similarly, such insights may concern questions of local culture history, span tens or even hundreds or thousands of years, or reflect more general and theoretical questions relating to the evolution of cultural systems. Archaeological sites can also supply information on past climates and vegetation patterns (e.g. through pollen grains), past fauna (e.g. through shell and bone remains) or inter-regional interaction through the geographical spread of ceramic traditions (e.g. through pottery sherds). In general, the scientific significance of sites increases as their potential information content increases. The archaeological significance of sites can be determined ‘according to timely and specific research questions on the one hand, and representativeness on the other’ (Bowdler, 1984:1). In terms of the former, detailed knowledge is required on the current state of play in academic archaeology – both in terms of local culture history and more general substantive, methodological and theoretical issues at the national and even international scales. Representativeness relates to the ability of a sample of sites from a particular area to represent as accurately as possible the range (and often frequency) of site types from a particular area (McMillan et al., 1977:32). As Lipe (1977:30) notes, ‘a representative sample is designed to represent a large population of items in terms of a small selection of such items with a minimum bias in the selection’. As a general rule the more rare a site, the greater its significance. It is in this sense that older sites tend to have greater significance given that older sites tend to be more rare due to the vagaries of time and decay (Coutts and Fullagar, 1982:61). However, an area exhibiting numerous similar (read common) sites can have considerable significance as it may provide a rare opportunity to investigate past land-use patterns. In this instance, the significance of the area is greater than the sum of its constituent sites (see also Bowdler, 1983:40). Furthermore, while a site type containing particular kinds of artefacts (e.g. a particular ceramic type) may be common in one, circumscribed area, at a groader geographical scale it may be rare (e.g. it may be largely restricted to this area). Such sites as a group may be assessed for significance, e.g. for its ability to reveal information about that area in relation to the broader region’s history. From a different perspective, representativeness also relates to maintaining the diversity of archaeological sites for future generations. This notion helps compensate for the biases inherent in academic research agendas that may ignore certain site types today but focus on these in the future (King et al., 1977:99). Examples of cultural heritage sites with very high scientific significance are Kuk (Western Highlands Province), one of the earliest known agricultural sites in the world; Kikiniu near Kopi (Gulf Province), an early village site with large quantities of traded ceramics; and OJP (Gulf Province), the only known Pleistocene site from the PNG southern lowlands. 4.5.12.3 Historical Significance A place has historical significance if it is associated either with significant person(s), event(s) or themes. As Kerr (1990:10) notes, the first two ‘may include incidents relating to exploration, settlement foundation, Aboriginal-European contact, disaster, religious experience, literary fame, technological innovation and notable discovery’. Historical significance may also include the ability of a place to be representative of major historical themes or cultural patterns from a particular historical period (Moratto and Kelly, 1978:4). As a general rule, the greater the degree of physical intactness of a site and its setting the greater its significance (Lennon, 1992:4). Examples of cultural heritage sites with very high historical significance are Samarai Island (Milne Bay Province), one of the first towns established by Europeans in PNG, and Dopima on the island of Goaribari (Gulf Province), where the early missionary James Chalmers was killed in 1901. B. David, B. Duncan, M. Leavesley 4-627 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.12.4 Educational and Economic Significance Cultural heritage sites may have important educational significance by providing opportunities for people (including community youths in indigenous learning contexts) to visit, examine and better appreciate the nature of these sites for themselves. Such opportunities not only have important or indeed profound social consequences in terms of maintaining a community’s identity, authenticity and sense of place (Lipe, 1984:6), but also can have significant economic consequences in terms of cultural tourism and a group’s ability to maintain itself economically (Hall and McArthur, 1993b). Cultural heritage sites also provide opportunities for the formal education sector. For example, the proximity of cultural heritage sites to teaching institutions is an important dimension of a site’s potential educational significance. In PNG, perhaps the very best examples of this has come from the involvement since the 1970s of University of PNG archaeology students in research-oriented excavations both on Motupore Island, to the south of Port Moresby, and at Boera, both of which are located within close proximity to the university’s Waigani campus. From another perspective, economic significance of sites is increasingly becoming an issue competing with alternative land-use activities (e.g. development). Although traditionally seen as mutually exclusive pursuits, cultural heritage preservation and economic development may work together. Best results occur where heritage issues are considered and accommodated for in the early stages of development planning (Rickard and Spearritt, 1991). An example of a cultural site with very high educational and economic significance is the Kokoda Trail (Northern Province). 4.5.12.5 Aesthetic Significance The aesthetic qualities of a cultural heritage place relate to the visual appeal, however subjective, of sites and their settings (Kerr, 1990:10). Despite the poorly defined nature of aesthetic significance, it remains one of the most important criteria for official registration of heritage sites in many parts of the world (e.g. Schapper, 1993). A famous example of a cultural site with very high aesthetic significance are the impressive terraces at Bobongara on the Huon Peninsula (Morobe Province), where cultural materials dating back to about 40,000 years ago have been found. 4.5.12.6 Significance of Recorded Cultural Heritage Sites First and foremost, it is important to point out that the general Boera-Manumanu area is a cultural heritage hot-spot in PNG. As the historically documented centre of the hiri trade, it was until very recently one of PNG’s foremost pottery-producing localities, each ethnographic and ancestral Motu village in particular representing a centre of craft industry and specialization. This has ensured a rich assemblage of archaeological sites in many locations within this broader area, the shallow marine and landscape immediately surrounding Boera in particular being of immense social, historical, scientific and educational significance and archaeological potential. The following discussions of significance are all founded on this clear and unambiguous understanding. Measures of significance for each of the cultural heritage sites recorded during the surveys are listed in Table 4.5.22. For each site, we make a separate assessment of significance for each of the five significance criteria outlined above. Our subsequent discussions of impacts on cultural heritage sites (section 4.5.13), and our recommendations based on those anticipated impacts (section 4.5.14), revolve around these measures of significance. 4.5.12.6.1 Levels of Significance Sites of Local Significance. All of the recorded cultural heritage sites are of at least low levels of significance at a local level, in the sense that all of the sites we recorded are important to local communities (social significance), and in one way or another contribute to our historical and scientific understanding of local culture (i.e., no site can be said to have no cultural heritage value at all). With B. David, B. Duncan, M. Leavesley 4-628 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project this proviso in mind, because all sites possess this minimal level of significance, we concentrate here on discussing higher (Provincial and National) levels of significance. Sites of Provincial Significance. 116 individual cultural heritage sites are of medium, high or very high Provincial significance. Most of these are of Provincial significance because of their social, scientific or historical and educational values. In all cases, old villages are Provincially important social sites because they enable the history (including occupation) of the broader region (and sometimes of entire clans) to be tracked back in time within the Central Province and beyond. For similar reasons, these same sites are often of Provincial scientific, historical and educational significance. Any site with good potential to reveal information of regional ceramic sequences will hold Provincial or greater significance. Furthermore, many WWII sites are Provincially important due to their historical, social and educational associations with those who served at the sites, and their associations with the defence of other regional centres around PNG. Sites of National Significance. 135 sites are identified as, or likely to be once more detailed assessment research is undertaken, of medium, high or very high National significance. The reasons why more sites are ranked of National rather than Provincial significance is, firstly, that a number of sites offer excellent, practical opportunities for University of PNG education and training (as has previously been the case for archaeological sites near Boera in the form of UPNG archaeology training programmes) (i.e. educational significance); and, secondly, that some WWII sites (such as crashed planes) hold greater prominence as national war graves than as provincially important sites. Some of these sites (e.g. AMG, AMH, JD2, CB10, CB11, CB12) are rated very highly because of their proven or likely archaeological worth as evidenced by eroding ceramics and dense cultural deposits capable of revealing historical information on southern PNG’s ancient history and broader trade relations prior to the coming of Europeans, but more research is needed at some of these sites before their significance can be properly assessed. Although highly and very highly significant sites appear to be clustered in this Boera-Lea Lea area (this is because this was a focal area for pottery manufacture and the centre of hiri trade), such sites with stratified deposits represent generally rare and at times unique opportunities to investigate various dimensions of a nation’s cultural heritage. Buria (CB10) is of further National significance because of its very high archaeological potential to reveal a unique snapshot of village life given that the site has remained virtually undisturbed since the village’s dramatic downfall and abandonment. Other village sites known from oral tradition (e.g. Aemakara, Daeroto, Dirora, Davage) are of great significance because they represent ancestral and origin villages for Western Motu and Koita generally, who now live so close to the nation’s capital Port Moresby and in so-doing have borne the brunt of colonial-period uran expansion (at the expense of many of their cultural sites). Sites CB6, CB7, CB8, ASM, Davage and Taubarau (among other, yetunrecorded sites) are of very high National historical, social and educational significance because they are directly associated with the first lagatoi story of Edai Siabo. Individually and together, these are major sites from which many Motu and Koita trace their ancestral roots, and because of their great significance as origin sites to one of PNG’s iconic indigenous social institutions (the hiri). These sites are a significant part of PNG’s National heritage (as evidenced by the use of the lagatoi symbol in many of Port Moresby’s and PNG’s public and private businesses and institutions). Similarly, the three sunken lagatoi sites are likely to be of National significance once their exact locations are found, for similar reasons concerning the iconic status of the hiri. Sites CB13, CB14 and CB15 are of potential National significance because they hold promise to contain submerged village remains (including wooden objects and structures). The Boera battery site is of National significance for its historical, social and political, and educational values because it is a location whereby Port Moresby was protected from invading forces during WWII, a defining episode in the history of the nation. Many of the plane crash and shipwreck sites (CB21-CB29) are of National significance because of their historical, social and political, and educational values and associations with WWII, and because of their significance as international war graves. B. David, B. Duncan, M. Leavesley 4-629 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project It is noted that while individual sites may each be assessed for their social and political, scientific, historical, educational and economic, and aesthetic significance at Local, Provincial and National levels of significance, sites may also attain various levels of significance because of the way they operate as a group. Therefore, while we discuss our reasonings for attributing certain levels of significance to individual sites below, we also in some cases identify groups of sites that together attain particular levels of significance (see section 4.5.12.6.3). It is important to point out at this stage that in some cases individual sites or groups of sites may potentially be of very high National scientific significance, without their actual level of significance being able to be properly determined until archaeological excavations have been undertaken (for it is only then that we would know their contents, which may or may not include unique, localized ceramic traditions that were traded broadly across southern PNG, or that could include archaeological materials from early phases of long-distance maritime trade preceding the kinds of pottery used in the ethnographic hiri trade). In all such cases, cultural heritage sites show good promise to contain cultural sequences capable of revealing considerable dimensions of cultural history for this part of PNG. Where a site has good potential to contain a particular significance value, it has been given such rating. Such sites must not be damaged in any way until appropriate archaeological sampling has been undertaken. The recommendations take these factors into account. 4.5.12.6.2 Individual Sites The individual sites of greatest importance are those ranked as having very high significance at the National level. These sites include: JD1-JD3, JD5, JD8-JD13, JD15, JDA17, ASM, CB6-CB8, CB10CB18, CB30, AADI, ABG, AMG, AMH, AMI, ANA, ANU, AOG, AOH, AOI, AOJ, AWL, Aemakara, Daeroto, Davage, Dirora and Taubarau. These sites are rated very highly because they represent places with the ability to furnish data reflecting cultural activity not otherwise available from other data sets, and/or because of their great significance as cultural ancestral places, or as WWII sites. This includes their potential to provide data that will contribute to archaeological research questions of both national and international interest. One particular group of these sites (JD1-JD3, JD5) also has extremely significant social value as they represent part of the location of the ethnographically documented village of Konekaru. This ancestral village has direct and immediate importance for the people of Papa today. These sites also have great potential to provide important educational opportunities for University of PNG archaeology, anthropology and cultural heritage management students. Section 4.5.12.6.1 already discusses the great significance of the sites associated with the legendary Edai Siabo (CB6, CB7, CB8, ASM, Davage, Taubarau), the Boera battery and the WWII plane and ship wreck sites (CB21-CB29). Forty-four cultural heritage sites are rated as of medium significance, and 24 of high significance. This is derived from the fact that although not all are dense sites, some are rich cultural deposits while others are unusually situated in the landscape in that they are located >1.0 km from the coast, an area in which pre-colonial period cultural activity has a generally relatively low archaeological signal (and an area through which the Koita came during the coast-ward migrations). For example, three of these sites – LNG 3, ML7 and ML9 – are large to very large in size and show very good archaeological promise, while JD17, ML1 and ML5 are identified as of medium significance because while they are located slightly more towards the coast than the other sites in this group, they possess stratified or probably stratified cultural deposits capable of revealing important information on the area’s cultural history. The sites rated as of relatively low significance were rated such either because they are located in a secondary context (i.e. they are disturbed cultural deposits), or they have a relatively small quantity and range of archaeological material with low potential for revealing information on the cultural past. B. David, B. Duncan, M. Leavesley 4-630 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.12.6.3 Site Complexes A site complex (listed as ‘SC’ in Table 4.5.22 and from here-on) is a group of discrete sites, usually located within a finite area that for archaeological purposes may be considered to represent interrelated activity areas. Within and near the Study Area, seven mutually exclusive site complexes have been identified. Site Complex 1. SC1 consists of sites LNG1, LNG2, LNG3, LNG4, ML1, ML4, ML5, ML7 and ML9. It is a series of low-density pottery scatters on a strip of land running N-S, located mostly >1 km inland broadly parallel to and on the seaward side of the Boera-Papa Road at the northern end of the Study Area. While individually the sites have limited National significance values, together they promise to reveal significant insights as to past use of the coastal hinterland where large archaeological sites and site complexes are seldom found. The low-gradient hillslope where SC1 is found evidences pronounced use sometime in the past, a geographical zone for which the archaeological signature is relatively weak. Therefore, this site complex is rated of moderate National but very high Local significance on the basis of the fact that collectively the sites have the potential to provide important information about cultural activity in an area removed from the immediate coastal strip (an area through which in ancient times the Koita migrated to the coast and through which more hinterland hamlets subsequently maintained contacts with coastal settlements; see section 4.5.7.1). Site Complex 2. SC2 consists of sites ML6, ML8, ML10, ML11, ML15, ML16, ML19 and JD7. It is a series of relatively small, low-density sites located along a low ridge that is the eastern bank of the mudflats south of Konekaru. While individually the sites are considered to have limited National significance, together they are of greater importance because they reflect apparently locally unique cultural activities (as evident by their spatially discrete location). Site Complex 3. SC3 consists of sites ML12, ML14, ML17, ML18, ML20 and ML21. It is a series of relatively smaller sites situated on the western flood-plain of Ruisasi Creek. While individually the sites are of limited National significance, together they are of greater importance because they reflect cultural activities on the periphery of an important archaeological village site (SC5) to the south. Indeed, it is yet to be determined whether SC3 is the northern extension of SC5. Sites and site complexes of this nature are significant because often they reflect cultural activities that for one reason or another were not undertaken at the centre of a settlement; that is, sites at the peripheries of villages can reveal significant and unique information about aspects of life not carried out in public spaces (e.g. sacred-secret events). Site Complex 4. SC4 consists of sites CB1, JD1, JD2, JD3, JD5 and ML3. It is a series of Nationally significant sites that together represent part of the former ethnographic village of Konekaru. While the JD1 is of very high significance in its own right, the associated sites only serve to increase Konekaru’s archaeological heritage value. Site Complex 5. SC5 consists of sites JD8, JD9, JD10, JD11, JD12, JD13, JD14, JD15 and JD16. It is a series of Nationally significant sites that make-up a former, previously undocumented archaeological village site to the south of the Study Area. While many of these sites are individually of very high significance in their own right, the close geographical association of the sites in a relatively small cluster only serves to increase the site complex’s heritage value. Additionally, it is yet to be determined whether this site complex is to be associated with SC3. Only excavation will ultimately determine this potential historical connection. Site Complex 6. Sites CB6, CB7, CB8 and ASM should be treated as part of an integrated story-line, associated with the traditional story of Edai Siabo, the first lagatoi and the origins of the ethnographic hiri journeys and trade. There will certainly be other sites relating to this story that should also be part of SC6 (e.g. Taubarua, Davage), but these have not yet been systematically recorded. These sites as a group are of National significance for their social, historical and educational values, as well as for B. David, B. Duncan, M. Leavesley 4-631 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project their scientific significance capable of informing us on the history and evolution of the hiri trade. As a unified set of sites they are an important example of an active traditional belief system. Site Complex 7. This consists of sites AHW, ANN, ANO and CB10 at Buria generally. This is an immensely important cultural heritage site to people at Papa and Lea Lea in particular, an ancestral village site that is the source of important oral traditions. The Buria area undoubtedly contains many other cultural heritage sites that have not yet been recorded (because this area lies a significant distance beyond our brief), but we list it here to alert the developers of its very high significance with cumulative impacts and potential future expansions in mind. Figure 4.5.32 highlights the locations of the five very highly significant site complexes within and adjacent to the site security fence area. Figure 4.5.32 Location of highly significant site complexes SC1 to SC5. SC6 is not shown because it consists of a series of non-contiguous cultural heritage sites, and SC7 because it is located a considerable distance to the north of the site security fence area. B. David, B. Duncan, M. Leavesley 4-632 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National Aesthetic Local Provincial National LNG1 L L L M M M L L L L L L L L L LNG2 L L L M M M L L L L L L L L L LNG3 L L L M M M L L L L L L L L L LNG4 L L L M M M L L L L L L L L L ML1 L L L M M M L L L L L L L L L ML2 L L L M L L L L L L L L L L L ML3 VH M L VH H H L L L VH H H L L L ML4 L L L H M M L L L L L L L L L ML5 L L L H M M L L L L L L L L L ML6 L L L M L L L L L L L L L L L ML7 L L L H M M L L L L L L L L L ML8 L L L M L L L L L L L L L L L ML9 L L L H M M L L L L L L L L L ML10 L L L M L L L L L L L L L L L ML11 L L L M L L L L L L L L L L L ML12 L L L H L L L L L L L L L L L ML13 L L L L L L L L L L L L L L L ML14 L L L H M L L L L L L L L L L ML15 L L L M L L L L L L L L L L L ML16 L L L M L L L L L L L L L L L ML17 L L L H M L L L L L L L L L L ML18 L L L H M L L L L L L L L L L ML19 L L L M L L L L L L L L L L L B. David, B. Duncan, M. Leavesley 4-633 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National Aesthetic Local Provincial National ML20 L L L M L L L L L L L L L L L ML21 L L L M L L L L L L L L L L L JD1 VH M L VH VH VH H M L VH H H L L L JD2 VH M L VH VH VH L L L VH H H L L L JD3 VH M L VH VH VH L L L VH H H L L L JD5 VH M L VH VH VH L L L VH H H L L L JD6 L L L M L L L L L L L L L L L JD7 L L L M L L L L L L L L L L L JD8 L L L VH VH VH L L L L L L L L L JD9 L L L VH VH VH L L L VH H H L L L JD10 L L L VH VH VH L L L VH H H L L L JD11 L L L VH VH VH L L L VH H H L L L JD12 L L L VH VH VH L L L VH H H L L L JD13 L L L VH VH VH L L L VH H H L L L JD14 L L L L L L L L L L L L L L L JD15 L L L VH VH VH L L L L L L L L L JD16 L L L M M L L L L L L L L L L JD17 L L L M M M L L L L L L L L L AAHL L L L L L L L L L L L L L L L AAHM L L L H H M L L L H M M L L L AAHN L L L H H M L L L H M M L L L AAHO L L L M L L L L L M M M L L L B. David, B. Duncan, M. Leavesley 4-634 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National Aesthetic Local Provincial National AAHP L L L M L L L L L M M M L L L AAHQ L L L L L L L L L L L L L L L AAHR M L L M M L L L L M M M L L L AAHS M L L M M L L L L M M M L L L AAHT L L L L L L L L L L L L L L L AAHU L L L M L L L L L M M M L L L AAHV L L L M L L L L L M M M L L L AAHW L L L L L L L L L L L L L L L AAHX L L L M M L L L L M M M L L L AAHY L L L L L L L L L L L L L L L AAHZ L L L L L L L L L L L L L L L AAIB L L L L L L L L L L L L L L L AAIC L L L L L L L L L L L L L L L AAID L L L L L L L L L L L L L L L AAIE L L L L L L L L L L L L L L L AAIF L L L L L L L L L L L L L L L AAIG L L L M M L L L L M M M L L L AAIH L L L L L L L L L L L L L L L AAII L L L L L L L L L L L L L L L AAIJ L L L M M L L L L M M M L L L AAIK L L L L L L L L L L L L L L L AAIL L L L L L L L L L L L L L L L B. David, B. Duncan, M. Leavesley 4-635 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National Aesthetic Local Provincial National AAIM L L L M M L L L L M M M L L L AAIN L L L M M M L L L M M M L L L AAIO L L L M M M L L L M M M L L L AAIP L L L H H M L L L H M M L L L AAIQ M L L H M M L L L H M M L L L AAIR L L L M M M L L L M M M L L L AAIS L L L L L L L L L L L L L L L AAIT M L L M M M L L L M M M L L L AAIU L L L M M M L L L M M M L L L AAIV L L L L L L L L L L L L L L L JDA17 H H VH L L L H H H M M M L L L CB1 VH H H VH H H M M M H H H M L L CB3 L L L L L L M M L M M L L L L CB4 H M M L L L H M H H H H L L L CB5 L L L L L L M M M L L L L L L CB6 VH VH VH L L L VH VH VH VH VH VH CB7 VH VH VH M M M VH VH VH VH VH VH CB8 VH VH VH M M M VH VH VH VH VH VH M M M ASM VH VH VH M M M VH VH VH VH VH VH M M M CB10 VH VH H VH VH H VH VH H VH VH VH CB11 VH VH H VH VH H VH VH H VH VH VH CB12 VH VH H VH VH H VH VH H VH VH VH CB13 VH H H VH VH H VH VH M VH VH VH M M M B. David, B. Duncan, M. Leavesley 4-636 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National Aesthetic Local Provincial National CB14 VH H H VH VH H VH VH M VH VH VH M M M CB15 VH VH VH VH VH H VH VH VH VH VH VH M M M CB16 VH VH VH VH VH VH VH VH VH VH VH VH CB17 VH VH VH VH VH VH VH VH VH VH VH VH CB18 VH VH VH VH VH VH VH VH VH VH VH VH CB19 VH VH H L L L VH VH H VH H H L L L CB20 VH VH M L L L H M L VH M M L L L CB21 M M M L L L H M H H H H L L L CB22 H M M L L L H M H H H H M M M CB23 M M M L L L H M H H H H L L L CB24 M M M L L L H M H H H H L L L CB25 H M M L L L H M H H H H L L L CB26 M M M L L L H M M M L L L L L CB27 M M M L L L H M M M L L L L L CB28 H M M L L L H M H H H H L L L CB29 M L M L L L H M M M L L L L L CB30 VH VH VH H H H VH VH VH VH VH VH H H H CB31 M L L M M M M M M M M M JD4 L L L L L L L L L M L L M L L AADI VH H H M L L VH VH VH AAGM L L L L L L L L L ABG VH VH VH VH VH VH VH VH VH VH VH VH ABH H M L M M M M M M M M M B. David, B. Duncan, M. Leavesley 4-637 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National ABI Scientific Local Provincial Historical National VH VH H H H M Educational and Economic Local Provincial National L L L Local Provincial National H H H H H H AEZ AFA AFB AFC VH VH H VH VH H H H M VH VH H AHW VH VH H VH VH H VH VH H VH VH M AHY L L L L L L L L L L L L AMG VH VH VH VH VH VH VH VH H VH VH VH AMH VH VH VH VH VH VH VH VH H VH VH VH AMI VH VH H VH VH H VH VH H VH VH VH ANA H H M VH VH VH H H M VH VH VH ANN VH VH H VH VH H VH VH H VH VH M ANO VH VH H VH VH H VH VH H VH VH M ANU VH VH VH VH VH VH VH VH H VH VH VH AOG H H M VH VH VH H H M VH VH VH AOH H H M VH VH VH M M M VH VH VH AOI H H M VH VH VH M M M VH VH VH AOJ H H M VH VH VH M M M VH VH VH AOK H H M M M M H H H AOL M M M L L L M M M AOM AOX APC B. David, B. Duncan, M. Leavesley 4-638 Aesthetic Local Provincial National Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National Aesthetic Local Provincial VH M National APF APG L L M L L L L L H L L ARD M L L H H M L L L H H H ARE M L L M M M L L L H M M ARF M L L M M M L L L H M M ARG M L L H M M L L L H M M ARH M L L H M M L L L H M M ARI M L L H H M L L L H M M ARJ M L L H H M L L L H H H ARK M L L H H M L L L H H H ARL M L L H H M L L L H H H ARM M L L H H M L L L H H H ARQ L L L H H M L L L H H H M L VH VH H L L L VH VH VH AWL Aema kara (oral traditio n site) VH VH H VH VH VH VH VH H VH VH H Daerot o (oral traditio n site) VH VH H VH VH VH VH VH H VH VH H B. David, B. Duncan, M. Leavesley 4-639 L Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National vDiror a (oral traditio n site) VH VH H VH VH VH VH VH H VH VH H Davag e (oral traditio n site) VH VH VH VH VH VH VH VH VH VH VH VH Konek aru village (oral traditio n site) VH M M H M M H H M VH H M Apau village (oral traditio n site) VH M M H M M H H M VH H M Tauba rau (oral traditio n site) VH VH VH H H H VH VH VH VH VH VH Boera freshw ater wells VH L L M M M H L L H M M B. David, B. Duncan, M. Leavesley 4-640 Aesthetic Local Provincial M M National M Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National Aesthetic Local Provincial National Sisal Farm/ Schim mer airstrip /Fairfa x Cattle Statio n M L L L L L M M M M M L L L L P-39 Airaco bra (Serial # 416945) plane crash site M M M L L L H M H M M M L L L P-38 Lockh eed Lightni ng plane crash site, Lea Lea airstrip M L L L L L M M M M M M L L L B. David, B. Duncan, M. Leavesley 4-641 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.22 Level of significance for the individual cultural heritage sites and site complexes recorded during the present surveys. (cont’d) Level of Significance Site # Social and Political Local Provincial National Scientific Local Provincial Historical National Educational and Economic Local Provincial National Local Provincial National Aesthetic Local Provincial National B-24 Conso lidated Libera tor plane crash site M M M L L L H M H H H H L L L SC1 L L L VH H M L L L VH H M L L L SC2 L L L H M M L L L H M M L L L SC3 L L L H M M L L L H M M L L L SC4 H M L VH VH VH L L L VH VH VH L L L SC5 L L L VH VH VH L L L VH VH VH L L L SC6 VH VH VH VH VH VH VH VH VH VH VH VH L L L SC7 VH VH H VH VH VH VH VH H VH VH VH L=Low, M=Medium, H=High and VH=Very High. SC= Site Complex. Significance levels are either actual (where sites have been assessed in the field), or likely or potential (where sites have been assessed from archives or oral traditions). In many cases where sites identified from the literature have not been relocated, it is not possible to assess aesthetic significance. In some cases of archaeological sites located at some distance from the proposed LNG Facilities site area, and where these sites are known from PNG National Museum and Art Gallery register records only, levels of significance cannot be determined where the details of site cultural contents are not presented in register records (in such cases, only the location of an archaeological site is known). B. David, B. Duncan, M. Leavesley 4-642 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 4.5.13 Impact Assessments This section discusses anticipated impacts of the proposed developments on the cultural heritage sites documented in the previous sections. From the onset, it is stressed that the territorial lands and waters of the Western Motu – and in particular the broader Boera region – are a cultural heritage hotspot, largely because this was a major centre of pottery manufacture; a major centre for both Motu and Koita settlement; and, according to legend, the origin location of the hiri trade. However, the area’s cultural heritage sites do not only consist of places known through oral history. They also include ancestral Motu and Koita sites that have been lost to social memory but that are still present as archaeological sites. These archaeological sites contain unique evidence on the history of Koita and Motu cultural practices: how Koita and Motu culture changed through time, including settlement locations, ceramic conventions, patterns of inter-regional interaction, inter-tribal alliances and conflicts, and trade relations. Such cultural dynamics have taken place in shifting locations of cultural activity, which today is evident in a large number of cultural heritage sites both within and neighbouring the proposed LNG Facilities site. Some of these cultural heritage sites are small, and some are very large and very rich in cultural deposits. Before anticipated impacts on these cultural sites are addressed, it is stressed that Motu and Koita (and, further to the east, Koairi as well) cultural sites have arguably suffered more damage than those of any other group in PNG, due to the footprint of Port Moresby and associated urban growth: the Motu and Koita live in the area of the nation’s capital, and as such their cultural heritage sites have been subject to ongoing damage and destruction through urban and associated developmental encroachments. The Boera-Papa area lies at the margins of such encroachments, and while to-date has remained largely unaffected by major developments, requires careful management in light of its proximity to the nation’s capital and its compounding development encroachments. It is worth emphasizing that cultural heritage sites are non-renewable cultural resources: once destroyed, they are gone forever. The Motu and Koita have already lost many sites over the years through urban developments, and it is critical that significant cultural heritage places remain for future generations to be able to access and engage with. The Boera area has a critical part to play in this, as shown by the presence of very significant cultural heritage sites as discussed in earlier sections of this report. In this section, anticipated impacts of the proposed LNG Facilities site developments are addressed, and recommendations by which to mitigate such impacts are made in section 4.5.14 to follow, with these concerns in mind. The proposed developments are likely to have impacts (in many cases of high magnitude) on numerous cultural heritage sites and materials within the Study Area and beyond. All the cultural heritage sites within the site security fence area will be subject to high impacts by the proposed developments, and associated infrastructures (e.g. access roads, accommodation blocks) as well as regional population growth caused by the proposed developments will also result in major impacts on cultural heritage sites in surrounding areas. The development of a management plan for the cultural heritage sites both within and near the site security fence area is thus of utmost importance (see section 4.5.14.3.1, Recommendation 4). Many of the archaeological sites within the site security fence area, and other sites nearby, contain sub-surface cultural materials. Unless adequately managed, for example by GIS plotting and mapping, the identification of buffer zones, and in some cases fencing off or systematic archaeological salvaging, some of the more significant cultural heritage materials and sites could be unacceptably destroyed or otherwise damaged through activities such as ground leveling, digging foundations for support facilities or fence posts, or through trench excavations, or through crushing in the passage of heavy vehicles or the positioning of containers, increased visitor traffic and the like. For this reason, and with due concern for the various values, levels and criteria of significance of the region’s cultural resources, it is imperative that an informed and systematic B. David, B. Duncan, M. Leavesley 4-643 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project management (including salvage) programme be instigated prior to commencement of the proposed LNG Facilities developments (see section 4.5.14 for recommendations). Likely impacts on each of the cultural heritage sites documented in this report are discussed so as to enable management recommendations to be made in section 4.5.14. As advised by Coffey Natural Systems, issues of impact are addressed through site-specific impact matrices aimed at assessing the valency, nature, duration, extent, magnitude and likelihood of impacts on cultural heritage sites (Table 4.5.23). These impact matrices are the same as those used in the PNG Gas Project, and are being adopted by all related archaeological projects (as advised by Coffey Natural Systems). The impact matrix is in each case presented as a set of choices identifying general impact concerns, as informed by the planned developments presented in this report and without consideration of the management recommendations made in section 4.5.14 (which are made as solutions to these anticipated impacts). The section 4.5.14 management recommendations are themselves designed to minimize the negative impacts presented in the section 4.5.13 impact matrices. The impact matrix cells have been infilled in orange to identify impact concerns for individual sites and site complexes. Because the field surveys have focused on areas within and adjacent to the site security fence area, and because also of the sensitivity of cultural sites to disturbance, generally the cultural heritage sites documented here are predicted to be negatively impacted in one way or another by the proposed LNG Facilities developments. However, and as will be discussed in the recommendations section 4.5.14 below, specific measures can be recommended to mitigate against developmental impacts by 1) maximizing information retrieval (salvaging; detailed recording) at some sites; and 2) avoiding damage (protection) to the most significant cultural sites where damage would be unacceptable. It is pointed out from the onset that the surveys by which such impact statements are made do not consist of a systematic 100 % survey coverage of the entire area of the proposed development (due to reasons outlined elsewhere in this report, largely to do with the discovery of UXOs and the ensuing restrictions made to the brief and methodology by Coffey Natural Systems). Furthermore, due to taphonomic processes of burial through time, some sites without surface exposure may occur under the ground and may thus not be detectable through archaeological surveys. The following statements are therefore based on the sample surveys undertaken, and on the sites that occur within the surveyed area. Yet cultural heritage surveys for impact studies cannot entirely be done through probabilistic surveys: a cultural heritage site could theoretically occur anywhere, and irrespective of how much of a surrounding area is covered by cultural heritage surveys, unpredicted and unknown sites could always be present in unsurveyed areas nearby. For this reason prior to commencement of developments it will be crucial to complete the cultural heritage surveys until 100% of the area is systematically covered by ground walking, all the sites discovered are recorded and management recommendations have been made for each site. This point will be further discussed in the section 4.5.14 recommendations. For the land area, both the recorded cultural heritage sites within the proposed site security fence area and those outside the fence are included. The reason for this is that impacts will not only be felt directly within the development area and the footprint of associated infrastructure (e.g., road upgrade and deviation), but also in surrounding areas (for a variety of reasons including, incremental growth, and indirect and cumulative effects). Similarly within the shallow marine environment, although the surveys have focused on a narrow corridor up to 700 m on either side of the proposed jetty and Marine Offshore Facilities at the time of the surveys, assessments of some other sites which fall outside these areas are also commented upon, as indirect and incremental impacts and dredging activities associated with the shipping channel and turning basins, along with any associated disturbance of the seabed, may occur. B. David, B. Duncan, M. Leavesley 4-644 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.23 Impact matrix for cultural heritage sites (courtesy of Coffey Natural Systems) Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure The following impact statements for the shallow marine environment are based on data (and field surveys) for locations which now appear to have been superseded by more recent plans for offshore facilities (see section 4.5.3 and section 4.5.5.1 for descriptions of the locations within which the cultural heritage site surveying team was asked to undertake fieldwork for this consultancy). 4.5.13.1 Cultural Heritage Sites Inside vs Sites Outside the Site Perimeter Fence Area (Land) and Study Area (Shallow Marine) Sites within the site security fence area are differentiated from sites outside the fence, and in the case of the shallow marine environment, the part inside the Study Area is differentiated from those outside the Study Area, because sites within the Study Area can be expected to be destroyed whereas those outside will in most cases be subject to lesser levels of disturbance. All the sites within the site security fence area, and some neighbouring sites, are expected to be subject to direct impacts. Therefore, qualitatively different management requirements will generally be required between these two sets of sites. 4.5.13.1.1 Direct Impacts: Cultural Heritage Sites Inside or Adjacent to the Site Security Fence Area ARD, ARG, ARH, ARI, ARJ, ARM, ML4, ML5, ML7, ML9, ML13, ML14, ML15, ML16, ML17, ML18, ML19, ML20, ML21, JD6; JD8, JD9, JD10, JD11, JD12, JD13, JD14, JD15, JD16 (SC5); JD17, AAHL, AAHM, AAHN, AAHO, AAHP, AAHQ, AAHR, AAHS, AAHT, AAHU, AAHV, AAHW, AAHX, AAHY, AAHZ, AAIB, AAIC, AAID, AAIE, AAIF, AAIG, AAIH, AAII, AAIJ, AAIK, AAIL, AAIM, AAIO, AAIR, AAIS, AAIT, AAIU, AAIV. These sites are all located within the site security fence area. Unless specific management recommendations are made for the protection of individual sites, each of these sites will almost certainly be significantly damaged or entirely destroyed during the proposed developments. B. David, B. Duncan, M. Leavesley 4-645 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.24 Impact matrix for cultural heritage sites ARD, ARG, ARH, ARI, ARJ, ARM, ML4, ML5, ML7, ML9, ML13, ML14, ML15, ML16, ML17, ML18, ML19, ML20, ML21, JD6; JD9, JD8, JD10, JD11, JD12, JD13, JD14, JD15, JD16 (SC5); JD17, AAHL, AAHM, AAHN, AAHO, AAHP, AAHQ, AAHR, AAHS, AAHT, AAHU, AAHV, AAHW, AAHX, AAHY, AAHZ, AAIB, AAIC, AAID, AAIE, AAIF, AAIG, AAIH, AAII, AAIJ, AAIK, AAIL, AAIM, AAIO, AAIR, AAIS, AAIT, AAIU, AAIV Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure LNG1, LNG2, LNG3, LNG4, ML12, JD7. These sites are all located directly on, or within a few metres, of the proposed site security fence area. Unless specific management recommendations are made for the protection of individual sites, each of these sites will almost certainly be damaged or destroyed during construction of the perimeter fence. Table 4.5.25 Impact matrix for cultural heritage sites LNG1, LNG2, LNG3, LNG4, ML12, JD7 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB3. This site, a WWII pipe bridge through the mangroves, is located either within (and at the southern end of) the site security fence area, or just outside of it. The exact location of the site was not recorded because at the time of the surveys it lay outside the Survey Area (and thus it lay outside the area of the study brief). Subsequently, when the location of the site security fence area was communicated to the cultural heritage site surveying team on 10 June 2008, it became apparent that this site would likely lie within or adjacent to the development area, and may thus be damaged or destroyed by the construction and use of the LNG Facilities. This site is of low significance and can be removed after it has been recorded. B. David, B. Duncan, M. Leavesley 4-646 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.26 Impact matrix for cultural heritage sites ML1, ML2, AAIN Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure Sisal Farm/Schimmer airstrip/Fairfax Cattle Station. The sisal farm, airstrip and cattle station lie directly within the bounds of the Study Area (and make up most of the land part of the Study Area closest to the coast). They will be directly affected by any works undertaken within the site security fence area. As the surveys of this area have not been completed, the exact extent and archaeological signatures of this site remain unknown. The impact matrix for this site is based on what has been seen so far, and probably represents an accurate representation for this site as a whole. Table 4.5.27 Impact matrix for Sisal Farm/Schimmer airstrip/Fairfax Cattle Station Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure 4.5.13.1.2 Indirect and Cumulative Impacts: Cultural Heritage Sites Outside the Site Security Fence Area ML1, ML2, AAIN. These sites are located 350-400 m to the north (ML1, ML2) and northwest (AAIN) of the site security fence area. They are likely to suffer damage (erosion, people picking up artefacts) from visitor traffic either from LNG Facilities workers (be they local villagers or outsiders) and/or villagers approaching the LNG Facilities site. B. David, B. Duncan, M. Leavesley 4-647 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.28 Impact matrix for cultural heritage sites ML1, ML2, AAIN Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure JD1, JD2, JD3, JD5, ML3 (Part of SC4); JD4; ML6, ML8, ML10, ML11, Konekaru Village. These sites are located 130-750 m to the north of the site security fence area, at and adjacent to Konekaru beach. Because they are associated with the closest beach to the LNG Facilities area, they are likely to suffer significant increases in visitation as a result of the proposed developments (by local villagers as well as visitors and people associated with the LNG Facilities in one way or another). This beach area will also very likely see increased marine access as canoes and dinghies traveling along this part of the coastline, immediately to the north of the proposed jetty facility, moor on the beach (whether due to bad weather or other reasons) as continued southward travel will be inhibited by an exclusion zone caused by the Marine Offshore Facility’s jetty. This area will thus require special management consideration. Table 4.5.29 Impact matrix for cultural heritage sites JD1, JD2, JD3, JD5, ML3 (part of SC4); JD4; ML6, ML8, ML10, ML11, Konekaru village Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure AAIP, AAIQ. These sites are located 250-500 m to the south of the site security fence area, on the edge of the mangroves. They are large stone artefact scatters, and as such are likely to suffer opportunistic collection or movement of stone artefacts from increased visitor foot traffic. B. David, B. Duncan, M. Leavesley 4-648 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.30 Impact matrix for cultural heritage sites AAIP, AAIQ Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure AHW, ANN, ANO, CB10 (SC7); CB11; CB12; CB7, CB8 and ASM (Some of the SC6 Sites); AADI, ABH, AMG, AMH, AMI, ANA, ANU, AOG, AWL, Ava Garau, Davage, Taubarau, Nemu, Daeroto, Dirora. These sites are located >2 km to the north (the SC7 sites); >1 km to the south (the SC6 sites); >2 km to the northwest (CB11, CB12); Dirora >2 km to the north-northeast; and >2 km to the south (the remaining sites, with the location of Daeroto being unknown) of the site security fence area, and are thus unlikely to suffer direct impacts from the proposed developments. However, because of their very high levels of significance, indirect impacts will almost certainly be felt, especially through two factors: 1) the tendency for development workers to visit interesting places they hear about while on (or off) the job (with or without guidance from local villagers); and 2) changing visitation, demographics, settlement patterns and access routes caused by developments among local populations resulting in increased access (and possibly housing/gardening pressures) by local peoples and their friends, relatives and incoming outsiders. These problems of cumulative and indirect impacts are very real ones that local peoples generally experience (and complain about) following development projects. Significant and renowned sites (such as the ones listed here) are most susceptible to this kind of impact. Table 4.5.31 Impact matrix for cultural heritage sites AHW, ANN, ANO, CB10 at Buria (SC7); CB11; CB12; CB7, CB8 and ASM (some of the SC6 sites); AADI, ABH, AMG, AMH, AMI, ANA, ANU, AOG, AWL, Ava Garau, Davage, Taubarau, Nemu, Daeroto, Dirora; Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure Aemakara. Aemakara is a very significant ancestral village site for the Koita. It is a renowned oral tradition site. Aemakara is extensive (it is estimated to be about 1.0 x 0.7 km in area), located 700 m to the south of the site security fence area, and contains rich archaeological deposits. It is also a high B. David, B. Duncan, M. Leavesley 4-649 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project point for the region, with excellent views in all directions. This combination of factors make Aemakara a likely target for people wanting a good view of the surrounding landscape. It is extremely likely that it will receive increased levels of visitor (foot, but also possibly vehicle) traffic as a result of the proposed developments and their ensuing changing regional demographics, including people (illegally) picking up artefacts. Its close proximity to the proposed developments also makes it amenable to encroaching developments. Table 4.5.32 Impact matrix for Aemakara Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB1. This site is likely to be indirectly impacted by proposed marine constructions as a result of wave action caused by ship traffic. The construction of the jetty along with any associated removal of the mangrove environment may also alter coastal geomorphological processes in the area, which will probably also subsequently affect the environmental stability of this site. The major effect to this site would likely be erosional. Table 4.5.33 Impact matrix for CB1 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB13, CB14, CB15. These are the old parts of Lea Lea, Papa and Boera villages that are now submerged underwater and contain buried archaeological remains. They are at some distance from the proposed developments, but may be impacted by any growth of the villages directly or indirectly resulting from the proposed developments (through new house constructions, changing erosional regimes and the like). They are also likely to be affected indirectly by erosion/sand accretion through the construction of the jetty and Marine Offshore Facility. B. David, B. Duncan, M. Leavesley 4-650 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.34 Impact matrix for cultural heritage sites CB13, CB14, CB15 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure AAGM, ABG, ABI, AEZ, AFA, AFB, AHY, AOH, AOI, AOJ, AOK, AOL, AOM, AOX, APC, APF, APG, ARE, ARF, ARK, ARL. These sites are located a short distance (200 m to >2 km) to the south of the proposed Site Perimeter Fence Area. They are particularly susceptible to people picking up artefacts such as potsherds while traveling between Port Moresby and the LNG152 Facility, or visiting Boera and other nearby places, or being taken to places by local people. Such (illegal) activities are very real and typical of growth areas. Table 4.5.35 Impact matrix for cultural heritage sites AAGM, ABG, ABI, AEZ, AFA, AFB, AHY, AOH, AOI, AOJ, AOK, AOL, AOM, AOX, APC, APF, APG, ARE, ARF, ARK, ARL Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure ARQ. This site is located a short distance (appears to be <100 m of road based on PNG National Museum and Art Gallery records) from the Boera-Papa Road, and will be easy access for people using the road to access the development area; this site is particularly susceptible to visitation from significantly increased demographics (both inside and outside the immediate LNG152 Facility), and thus foot traffic. It is susceptible to erosion from foot traffic and from people poking around. B. David, B. Duncan, M. Leavesley 4-651 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.36 Impact matrix for cultural heritage site ARQ Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB16, CB17, CB18. We are uncertain as to the likelihood of these sites being damaged as their current locations are unknown, but it should be noted that these lagatoi wreck sites would almost certainly be of National (and international) significance if discovered. The proposed Pipeline development outside of the Caution Bay area may impact upon these or other previously undiscovered shipwrecks (from both the pre-European contact and contact periods) or aircraft crash sites. Table 4.5.37 Impact matrix for CB16, CB17, CB18 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB30. This is a major WWII facility site of National significance. It is located >2 km to the south of the Site Perimeter Fence Area, and approximately 2.5 km from the Boera-Papa road. It is a well-known WWII site, and of relatively easy access. It is one of the most intact and undisturbed WWII sites in the country, and may also contain archaeological remains of prehistoric occupation. It must not be disturbed by the proposed developments, but it’s high-quality visibility (as a WWII complex in a good state of preservation) means that it is very susceptible to increased visitation from LNG152 Facility personnel and from indirect and cumulative local and regional traffic caused by the developments (including changing regional demographics). Management considerations will need to be given to this site, despite its distance from the Site Perimeter Fence Area. B. David, B. Duncan, M. Leavesley 4-652 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.38 Impact matrix for CB30 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB4. Further work is required to positively identify the exact location of this site. The site lies outside the Study Area of the study brief, but seems to lie very close to the site security fence area, and may thus be subject to impacts from the poposed developments. From oral accounts by local people, this site appears to lie shortly to the south of Vaihua River amongst the mangroves. This WWII plane crash site (of a P-39 Airacobra) is of high significance, and any disturbance may require the permission of the Australian or US governments as well as the PNG government. Until the exact location of this plane wreck is identified in the field, management recommendations cannot be adequately made (see section 4.5.14 for recommendations). Table 4.5.39 Impact matrix for CB4 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB6, CB22, CB23, CB24, CB25, CB26, CB27, CB28, CB29, P-39 Airacobra (Serial # 41-6945) Plane Crash Site; B-24 Consolidated Liberator Plane Crash Site. The construction of the proposed shipping channel may effect at least one aircraft (CB23/CB24) and possibly more, some of whose locations and occurrences may not be presently known (historical documentary records indicate that other aircraft may have crashed or disappeared in this general area; other previously unreported sites may exist in the broader area). Furthermore, as the present planned route for the Pipeline is unknown to the cultural heritage site surveying team, it is also possible (but currently uncertain) that other aircraft crash and shipwreck sites at Hidiha (CB22), Bavo (CB28, CB29), and other reefs inside Caution Bay (CB25, CB26, CB27) may be affected directly or indirectly by the installation of the proposed Pipeline through construction, dredging or localised scouring of the seabed. Similarly, CB6, a site of very high National significance associated with the first lagatoi story, is also located at Hidiha Island, and could be adversely affected by any plans to construct a pipeline across the reef in this B. David, B. Duncan, M. Leavesley 4-653 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project area. CB6 must not be damaged, and concerted efforts should be made to accurately located and map all of these sites. Some of the WWII wreck sites are likely to be international war graves. Table 4.5.40 Impact matrix for CB6, CB22, CB23, CB24, CB25, CB26, CB27, CB28, CB29, P-39 Airacobra (Serial # 41-6945) plane crash site; B-24 Consolidated Liberator plane crash site Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure JDA17. This is probably the location of a WWII plane crash. The site was recorded opportunistically during the surveys, although it lay beyond the area of the study brief. The site lies 2.2 km to the southeast of the site security fence area. It is likely to be subjected to occasional visitation as a result of the proposed developments, due to the significant anticipated increases in population and traffic across the area. Table 4.5.41 Impact matrix for JDA17 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB5. It is uncertain as to the likelihood of damage at this site as its exact location has not been ascertained, but it may lie close enough to the proposed development to suffer direct or indirect damage. As it has not been ascertained whether the cable is still being used, neither can its level of significance or impact be determined. Although the UNESCO maritime convention (UNESCO 2001) does not consider submarine cables under cultural heritage, the cable and its location would probably be of archaeological significance due to its associations with PNG/Australian WWII defences. The location of where the cable came ashore would be of historical significance due to its associations with the first telegraph cable between PNG and Australia. However, its location is of National historical significance as the first communication cable between Australia and PNG. B. David, B. Duncan, M. Leavesley 4-654 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.42 Impact matrix for CB5 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure Apau Village. Members of the cultural heritage site surveying team have been told by a local Boera villager that the ancient village site of Apau is located approximately 50-100 m south Davage, near Boera, which is the building site of Edai Siabo’s first lagatoi. Apau village was later replaced by Boera village. It has not been field-recorded, because this area lay outside the area specified in the study brief during the fieldwork. Therefore only the place name has been recorded, following its mention during discussions with local villagers. This site should be systematically mapped prior to commencement of developments. It is unlikely to be impacted by the proposed developments, but a final decision on this, and on its management, will need to await its field mapping. Table 4.5.43 Impact matrix for Apau village Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure Boera Freshwater Wells. The Boera freshwater wells have not been field-located during the cultural heritage surveys, because this area lay outside the area specified in the study brief during the cultural heritage fieldwork. Therefore only their existence has been recorded, as a cultural heritage place of historical standing following their mention during discussions with local villagers. This site should be systematically mapped prior to commencement of developments. It is likely to be impacted by the proposed developments, as a result of growing populations and ensuing requirements for greater freshwater access by neighbouring villagers. However, the exact nature of impacts on this site are unknown to as the cultural heritage site surveying team has not participated in the broader social study and the implications of social impacts on land use that are assessed in the project Social Impact Assessment. B. David, B. Duncan, M. Leavesley 4-655 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Table 4.5.44 Impact matrix for Boera freshwater wells Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure CB21. According to local oral traditions, this WWII plane crash site is located on the mud bank at the southern end of Lea Lea village, and is probably the Japanese Zero (A6M2 Model 21 Zero – Manufacturer # 3537) which is described historically (see section 4.5.10.1.1 above; Table 4.5.6) as having crashed at Boera. Table 4.5.45 Impact matrix for CB21 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure P-38 Lockheed Lightning Plane Crash Site, Lea Lea Airstrip. This site is located a considerable distance from the proposed LNG Facilities site and is unlikely to be impacted by it. The crashed plane and all its parts have already been removed to the DoMH Museum at Boroko, Port Moresby. Table 4.5.46 Impact matrix for P-38 Lockheed Lightning plane crash site, Lea Lea airstrip Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure B. David, B. Duncan, M. Leavesley 4-656 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project CB19, CB20, CB31 From current plans available to the cultural heritage site surveying team, it appears that sites CB19, CB20 and CB31 will probably not be affected by the installation of the Pipeline or LNG Facilities as presently proposed. CB31 is located some distance from the site security fence area and is therefore unlikely to be impacted by the proposed developments. CB19 and CB20 are fixed ‘instillations’ (memorial and tree) located in the middle of Boera and Lea Lea villages respectively, and are unlikely to be affected by developments associated with the proposed developments. Because of the above unlikelyhood of impacts on these sites, the impact matrix is left blank. Table 4.5.47 Impact matrix for CB19, CB20, CB31 Valence Nature Duration Extent Magnitude Likelihood Positive Direct FEED Localised High Uncertain Negative Indirect Construction Regional Medium Probable Cumulative Operation National Low Confident Closure 4.5.14 Recommendations The cultural heritage sites surveys undertaken in the present report involved limited recording of individual sites rather than their detailed mapping and detailed investigations. Based on the number, distribution, significance and likely impacts of the proposed onshore and offshore LNG Facilities developments on the cultural heritage sites within the Study Area, as presented in previous sections of this report, a number of recommendations are made to mitigate against the loss of cultural heritage sites, and the information these may hold, as a result of the proposed developments. These recommendations will need to be addressed prior to the construction of the proposed LNG Facilities. As they relate directly to the impacts or potential impacts of the proposed developments on cultural heritage sites, the costs of meeting these recommendations remain the responsibility of the developers. This section of the report (the section 4.5.14 recommendations) is made by the following co-ordinated group of cultural heritage personnel: Bruno David, Nick Araho, Jeremy Ash, John Dop, Brad Duncan, Alexandra Gartrell, Alois Kuaso, Matthew Leavesley and John Muke. 4.5.14.1 General Principles Recommendation 1. That any further cultural heritage work should involve the PNG National Museum and Art Gallery working in conjunction with professional consultant archaeologists. The rationale behind this is that the Museum is required to administer the National Cultural Property (Preservation) Act, with the reporting of cultural places and objects to the Museum being a requirement of the Act. Furthermore, the Museum is best positioned to properly assess notions of National heritage significance as they apply to PNG, being a national representative body for cultural heritage matters in this country. Second, the Museum is also responsible for the administration of the War Surplus Material Act, which deals predominantly with historical defence sites, and as such is similarly placed to assess the significance of these sites. Third, by virtue of the proposed LNG Facilities developments being located in PNG, it is appropriate that PNG archaeologists are formally involved in all stages of assessing national cultural heritage matters, including sites and objects identified during surveys. It is also recommended that, in future, the National Cultural Commission (NCC) should also be involved in cultural heritage studies that involve cultural heritage sites and/or B. David, B. Duncan, M. Leavesley 4-657 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project objects (because they are the PNG government organisation responsible for cultural development in PNG, which is relevant to the present cultural heritage work because oral tradition sites are involved). 4.5.14.2 Completion of Cultural Heritage Surveys 4.5.14.2.1 Land and Shallow Marine Environment within the Site Perimeter Fence Area and Associated Jetty Area Recommendation 2. It is imperative that a comprehensive (100% aerial coverage) archaeological site survey be undertaken of all footprint areas of the LNG Facilities site components – for both construction and operations – located within the site security fence area boundary before any developments proceed. Similarly, the entire shallow marine area to be developed, including shipping turning areas, should be systematically surveyed (including side-scan sonar surveys; see also Recommendations 3 and 4) for cultural heritage sites, with WWII plane and shipwrecks particularly in mind (a number of missing plane crashes occur in the general area, some of which may be international war graves). 4.5.14.2.2 Area Outside the Site Security Fence Area and Associated Jetty Area Recommendation 3. Any associated infrastructure areas, such as lay-down areas, access roads, helipads and the like, should be systematically surveyed for cultural heritage sites prior to developments, and appropriate management plans for cultural heritage sites thereby discovered drawn-up. 4.5.14.3 Management Plan, Mapping, Avoidance and Salvage 4.5.14.3.1 General Management Plan Recommendation 4. Given the high density of cultural heritage sites (archaeological and oral tradition) within and surrounding the proposed developments, and given in particular the presence of numerous sites of very high National significance outside but a short distance away from the site security fence area and associated shallow marine facilities, together with anticipated demographic and settlement growths, we strongly recommend that systematic cultural heritage surveys, and a management plan for cultural heritage sites, be undertaken for the entire area from Boera to Papa, from the coast inland to the foothills. This management plan should consider both indigenous sites and WWII sites. It is noted that as the LNG Facilities work proceeds, it will inevitably draw more human traffic into the region. Such new-comers and/or visitors may unwittingly visit cultural sites against the wishes of the local residents or against co-venture participant rules and protocols (such activities are routinely documented from other comparable operations in PNG and elsewhere; at times people move or even illegally remove artefacts); in some cases, people who have come to the area as a result of the LNG Facilities developments will likely visit sites as local villager guests. In order to mitigate against such impacts on nearby cultural heritage sites, we strongly recommend that systematic cultural heritage surveys be undertaken in areas beyond the scope of this report, and that these should include all the known important cultural heritage sites within a 3 km buffer zone around the site security fence area. The results of these surveys should be incorporated in the regional cultural heritage sites management plan. 4.5.14.3.2 Geomorphological Modelling Recommendation 5. The LNG Facilities site area may be subject to long-shore drift which in turn may adversely affect coastal geomorphological dynamics. It is possible that the introduction of new maritime infrastructure on the shoreline could have adverse effects on the stability of the sites at Konekaru beach, the aircraft wrecks at Lea Lea and Vaihua, any remains of stilt village sites at Lea Lea and Papa, and most significantly the traditional lagatoi canoe sites at Boera. The coastal sites CB6, CB8, CB9, CB13, CB14, CB15 and CB16 which lie outside the Study Area, may also be affected. All of these sites are considered of potential National cultural heritage significance, and in particular sites CB13, CB14, CB15 and CB16 have the potential for highly intact archaeological B. David, B. Duncan, M. Leavesley 4-658 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project deposits. This equally applies to any aircraft wreckage sites which are known or which may exist in the area, particularly CB21, and also to the Missionary Tree (CB20) at Lea Lea which is situated close to the water’s edge in a low-lying area. It is therefore recommended that the potential effects of the installation of the proposed Marine Offshore Facility and jetty on local known and potential archaeological sites are investigated, with the establishment of management recommendations in mind. Mitigative works should also be considered to retain the current foreshore edge and retard erosion around the villages of Papa, Boera and Lea Lea (possibly in the form of a seawall and groynes if deemed appropriate) if project developments are predicted or shown to impact on these sites. 4.5.14.3.3 Sites Requiring Systematic Recording and Mapping Recommendation 6. A number of very significant ancestral cultural heritage sites of oral tradition occur outside the site security fence area, but close enough to be subject to low to medium level indirect and cumulative impacts. They include: B10 (Buria), Aemakara, Apau Village, Dirora, CB11 (Darebo), CB12 (Dori Hill), Daeroto, four sites associated with Edai Siabo’s first lagatoi story (CB6, CB7, CB8, ASM), and three sunken lagatoi (CB16, CB17 and CB18; see also Recommendation 8); these sites are of very high cultural significance to the people of Lea Lea, Papa, Boera and/or Porebada. We recommend that these sites be systematically recorded and mapped prior to LNG Facilities developments proceeding, and should involve a social anthropologist to record the oral traditions associated with them, with their longer-term management (including monitoring of impacts) in mind. Any other associated works planned for the area (e.g. offshore pipeline routes, inland traffic routes or other associated infrastructure facilities) should consider these sites at the planning phase and be routed to avoid any negative impacts to them. It is critical that no developments directly relating to the LNG Facilities take place at these sites. These sites were all recorded opportunistically while doing the work for the present report. However, oral tradition or archaeological sites occurring outside the area of the study brief (centred on the Survey Focus Area) have not been systematically recorded. There are likely to be many other significant sites of oral tradition, and archaeological sites, in the broader area and these should also be systematically recorded (see Recommendation 4). Recommendation 7. The area within and immediately surrounding the site security fence area contains numerous locations with Koita and/or Motu language names. Once the proposed developments take place, many of these places will almost certainly disappear from social memory, as the land will be significantly modified and many of its features will disappear. We therefore recommend that a social anthropologist record the Koita and Motu place names and their associated oral traditions within and immediately surrounding the site security fence area prior to construction. Recommendation 8. The discovery of any sunken lagatoi would be of National significance to PNG, and of likely international significance to the archaeological community. It is recommended that if any lagatoi remains are discovered during any works for the proposed developments, the site including a 300 m buffer zone around it should be protected from developments. This includes the three known lagatoi wreck sites CB16, CB17 and CB18, whose precise locations have not yet been identified. In order to avoid damage to these sites, and giving due thoughts to their general proximity to the proposed developments, we here also recommend that the precise locations and significance of CB16, CB17 and CB18 be recorded in the field, with the involvement of a social anthropologist to document their cultural details. Recommendation 9. A number of archaeological sites recorded over 20 years ago and already on the PNG National Museum and Art Gallery national site register occur within the site security fence area (sites ARD, ARG, ARH, ARI, ARJ, ARM). It is uncertain whether or not some of these sites are geographically contiguous with four of the sites newly recorded for the present report (in particular, sites AAIM, AAIO, AAIP, AAIQ). It is recommended that these two sets of site codes be cross-checked in the field to determine whether or not any of the sites have been recorded twice (e.g. do ARE = AAIP; ARG = AAIQ; ARJ = AAIO), once under earlier PNG National Museum and Art Gallery/University of PNG surveys, and a second time during the present surveys. B. David, B. Duncan, M. Leavesley 4-659 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Recommendation 10. The Boera battery (CB30) to the north of Boera village is considered one of the most intact and undisturbed WWII sites in the country and may also contain remains of pre-colonial period occupation. The site is likely to incur increased levels of visitation as a result of the proposed developments. In light of such ancitipated increases, systematic archaeological mapping of this site by a professional surveyor working together with an archaeologist is recommended. Recommendation 11. The anomaly identified by side-scan sonar (ID10) should be inspected by divers prior to construction, and management recommendations made accordingly, as this anomaly was not inspected during the fieldwork. 4.5.14.3.4 Sites where Disturbance is to be Avoided Recommendation 12. Konekaru (CB1, JD1, JD2, JD3, JD5, ML3, including the beach area; SC4), an ancient village site with high cultural heritage values located in the northwest corner of the LNG Facilities site lease area and a short distance outside the site security fence area, must be avoided. This is the site of a former village of the ancestors of the residents of Papa, and a significant ancestral cultural heritage place. Consequently, this site has direct and immediate social links and high social heritage value for the contemporary residents of Papa. Recommendation 13. The ancestral oral tradition sites near Boera – including Davage, Taubarau, Ava Garau – are of the highest level of cultural heritage significance. The archaeological sites associated with and adjacent to these oral tradition sites (including AFA, AMG, AMH, AMI, ANA, ANU) are also of very high cultural heritage significance. It is critical that this entire area remains undisturbed by the proposed LNG Facilities development because of its very high National level cultural heritage significance. It is strongly recommend that the area which contains these sites from Boera village north and northwesterly-ward to AGD66 Grid Reference 0500800 E 8963700 N, plus a 1 km-wide buffer zone around it (including both the land and marine environments), is avoided by the LNG Facilities development (including associated infrastructure areas such as lay-down areas, access roads, shipping routes and turn-around areas, etc.). Recommendation 14. It is critical that the very significant ancestral oral tradition sites of Buria (B10), Aemakara, Dirora, Darebo (CB11), Dori Hill (CB12) and Daeroto are avoided by the proposed developments. No access roads, lay-down areas or other infrastructural developments for the project are to be undertaken within at least 500 m of any of these sites. Recommendation 15. Cultural heritage sites CB6, CB7, CB8 and ASM are very significant oral tradition sites associated with the first lagatoi story. It is critical that damage to these sites is avoided, and any developments that could come within 300 m of these sites should be reconsidered. The exact location of CB6 near Hidiha Island is as yet undetermined, but the site may be affected by any subsequent developments if this location is not sufficiently considered in the future. It is therefore strongly recommended that further cultural heritage survey work be undertaken to document the precise location and significance of CB6, and that a social anthropologist should be engaged to record the oral traditions surrounding these sites and any other associated cultural heritage sites. Recommendation 16. Cultural heritage sites CB16, CB17 and CB18 are sunken lagatoi wreck sites known through oral traditions; they are of great cultural heritage significance. Damage to these sites should be avoided at all costs, and any planned developments within 300 m of these sites should be reconsidered. 4.5.14.3.5 Sites Requiring Archaeological Salvaging Prior to Construction A number of archaeological sites hold good to very high potential to reveal unique information about the area’s and broader region’s history. These sites should each be salvaged prior to disturbance by 2 the proposed developments. The larger and denser archaeological sites (>1000 m ) should each undergo extensive salvaging; the smaller and less dense stratified sites should each undergo smallerscale salvaging. In all cases, archaeological excavation methods should follow international best B. David, B. Duncan, M. Leavesley 4-660 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project 2 practice methods, including systematic excavation of ≤1 m Excavation Squares (sometimes as single squares, sometimes in sets of contiguous or adjacent squares, depending on the specific needs of individual site and site complex salvaging programmes); excavation in fine-grained (usually <3 cm thick) Excavation Units within Stratigraphic Units (to enable appropriate differentiation of temporal units); sub-division of stratigraphic sub-units to enable differentiation of separate depositional/erosional features; systematic 3-dimensional plotting of significant in situ objects and charcoal for AMS radiocarbon dating; sieving in 3mm mesh or less; radiocarbon dating of each excavation to allow meaningful results to be presented; complete sorting of all excavated materials; laboratory analysis of excavated materials to a sufficient extent to allow a complete and systematic reporting of site contents to the PNG National Museum and Art Gallery. We stress that these minimal salvaging procedures are of utmost importance and necessary prior to construction because we recommend that the archaeological sites involved can then in most cases be damaged or destroyed (and in the other cases, sites may suffer some degree of indirect and incremental impact as a result of the developments). The salvaged data will therefore be the only systematic information ever available on the destroyed sites after construction proceeds. We recommend that small and sparse sites lacking stratified deposits do not require salvaging. Recommendation 17. Archaeological deposits at SC4 (the ancient village site of Konekaru) centre on 2 JD1 and CB1 should be test-sampled (i.e. a 1 m square be excavated at each site) by professional archaeological means (see also Recommendation 12 above). While we strongly recommend that this site not be disturbed by the proposed developments, it will almost certainly be subject to some degree of impact as a result of increased human traffic caused by the construction of the LNG Facilities. Because sites JD1 and CB1 in particular are likely to have large sub-surface archaeological deposits of National cultural heritage significance, and in light of their close proximity to the proposed LNG Facilities, additional to Recommendation 12, we also recommend limited archaeological salvaging of JD1 and CB1. Recommendation 18. SC5, consisting of JD8, JD9, JD10, JD11, JD12, JD13, JD14, JD15 and JD16, is a complex of sites representing an ancient village with very high cultural heritage value. It has substantial archaeological deposits spread across a large area extending from the western bank of Vaihua (Ruisasi) Creek (in the direction of SC3), with some sites being themselves extensive as individual sites. These sites are located within the site security fence area. We strongly recommend that SC5 is either avoided or extensive salvage be undertaken prior to construction of the LNG Facilities. This includes archaeological investigation of each individual site. Recommendation 19. Archaeological deposits at SC4 (the ancient village site of Konekaru) centre on 2 JD1 and CB1 should be test-sampled (i.e. a 1 m square be excavated at each site) by professional archaeological means (see Recommendation 14 above). While we strongly recommend that this site not be disturbed by the proposed developments, it will almost certainly be subject to some degree of impact as a result of increased human traffic caused by the construction of the LNG152 Facility. Because sites JD1 and CB1 in particular are likely to have very significant sub-surface archaeological deposits of National significance, and in light of their close proximity to the proposed LNG152 Facility, additional to Recommendation 14 we thus also recommend limited archaeological salvaging of JD1 and CB1. Recommendation 19. SC1 is a series of low-density cultural activity areas located immediately outside but close (20-300 m) to the site security fence area, and therefore likely to incur visitation impacts (from human visitation, foot traffic and vehicle traffic during construction and operation) as a result of project developments. It contains a number of sites (LNG1, LNG2, LNG3, LNG4, ML1) situated approximately 600 m to 1.5 km inland from the coast on low gradient seaward facing slopes. It is the only site complex of this type recorded in the study area and provides valuable data in regards to the use of this area. The archaeological deposits are spread across an extensive area. We strongly recommend that the SC1 sites be subjected to small-scale archaeological excavation prior to B. David, B. Duncan, M. Leavesley 4-661 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project construction in order to obtain sample information pertaining to the cultural activities undertaken in this area. Recommendation 20. SC3 (ML14, ML17, ML18, ML20, ML21) is a series of apparently discrete cultural activity areas located on the western bank of Vaihua (Ruisasi) Creek within the site security fence area. Although many of the individual sites are relatively small, they cluster in such a way to suggest that they may be a northern extension of the highly significant SC5 archaeological village complex (as outlined in Recommendation 18). We strongly recommend that each site within SC3 is subjected to small-scale archaeological excavation prior to construction in order to determine their relationships to SC5, as well as to obtain historical information pertaining to the cultural activities undertaken in this area. Recommendation 21. The ancient Koita village site of Aemakara contains rich archaeological deposits which remain largely unrecorded. Aemakara lies about 700 m to the south of the site security fence area. This very highly significant cultural heritage site is likely to be impacted by increased visitor 2 traffic as a result of the proposed developments. It is recommended that a small (1-3 m ) archaeological excavation be undertaken at this site to determine its precise archaeological significance, with long-term management in mind. Recommendation 22. In addition to the sites already listed for salvaging in previous recommendations, it is strongly recommended that the following archaeological sites containing (or likely to contain) stratified deposits each undergo limited systematic archaeological salvaging: ARD, ARG, ARH, ARI, ARJ, ARM, AAHM, AAHN, AAHO, AAHP, AAHR, AAHS, AAHU, AAHV, AAHX, AAIB, AAIC, AAIG, AAIJ, AAIM, AAIO, AAIR, AAIT, AAIU, ML4, ML5, ML7, ML9, ML15, ML16, JD6, JD8, JD9, JD10, JD11, JD12, JD13, JD14, JD15, JD16, JD17. These sites are all located within the site security fence area, and will thus almost certainly be extensively disturbed or destroyed as a result of the proposed developments. Of these, we recommend that prior to construction the small 2 archaeological sites with less than 25 m stratified cultural deposit undergo small-scale salvaging 2 2 2 (excavation areas = 1 m at each site); that 1-3 m are salvaged at each of the sites with 25-100 m 2 2 stratified cultural deposit; and that 1-25 m are salvaged at each of the sites with >100 m stratified cultural deposit. The final decision as to exactly how much of each site is to be sampled should be made by the cultural heritage team prior to and during archaeological excavation (such decisions are best refined as subsurface deposits are revealed during the process of excavation). In all cases, this would mean that a small proportion of the cultural items from each site is systematically sampled prior to site disturbance or destruction. Recommendation 23. All stratified archaeological sites found within the site security fence area during the completion of the systematic (100 %) cultural heritage sites survey, and in associated infrastructure areas (see Recommendations 2 and 3), should be sample-excavated in the same way as for the sites listed in Recommendation 22 above prior to construction. Recommendation 24. We recommend that the archaeological salvage work at all times involves 1), professional archaeologists working collaboratively with 2), appropriate clan/section village representatives and 3), archaeology or history students from the University of PNG. Such collaboration would serve the dual purpose of undertaking the necessary salvage work while at the same time meeting best practice standards of community participation and training as social responsibility initiatives. 4.5.14.4 Monitoring Recommendation 25. The very significant cultural heritage sites identified in Recommendation 6 will likely be subjected to indirect and cumulative impacts by the proposed developments. We recommend that these sites be periodically monitored by cultural heritage officers together with appropriate traditional clan representatives nominated by senior community members to assess such impacts and, B. David, B. Duncan, M. Leavesley 4-662 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project if necessary, to make new recommendations to address management issues that may result from the LNG Facilities developments. Recommendation 26. The cultural heritage sites reported in the present report have aimed to include all of the sites observable from the ground surface within the surveyed areas, and we have strongly recommended that all the areas not yet systematically surveyed within the site security fence area and associated infrastructure areas should be 100 % surveyed prior to construction (Recommendations 2 and 3). Furthermore, we have also recommended that all stratified archaeological sites within the site security fence area undergo systematic archaeological salvage sampling (Recommendations 18, 20, 22, 23). These systematic surveys notwithstanding, some archaeological objects, and some archaeological sites within the site security fence area, are likely to have no surface expression today, being buried at various depths below the ground. We anticipate that such archaeological materials will be exposed during construction of the LNG Facilities, although large sites are unlikely to appear because such large sites are likely to contain some degree of surface exposure (e.g. through erosion, animal burrows bringing sub-surface deposits to the surface and the like) by which they would have been identified during the course of the surveys. To cover for such possibilities of unknown archaeological sites being uncovered during the course of construction, we recommend that an archaeologist be employed to monitor clearance and sub-surface disturbance activities during construction of the LNG Facilities and associated infrastructure. In those cases where archaeological materials are exposed during construction, we recommend that the monitoring archaeologist flag the site’s location on the ground, and undertake rapid archaeological salvaging work. Such salvage work should not unduly delay construction activity, but be undertaken promptly at the first opportunity following site discovery. Recommendation 27. In any project development footprint area where cultural material is discovered on the seabed, it is recommended that a maritime archaeological survey of the seabed using divers/remotely operated vehicles be undertaken (dependent on water depth) for recording and if deemed necessary for rapid salvaging or protection purposes prior to the recommencement of works. Recommendation 28. Where dredging works are undertaken, it is recommended that a sample of dredge spoil be taken every hour to investigate the presence of submerged prehistoric artefacts. A trained cultural heritage officer should be on-board with this aim in mind (this would not have to be a fully trained archaeologist, but could be a community officer specially trained with this aim in mind). This procedure follows best-practice protocols, such as those currently being undertaken in the UK to detect ancient prehistoric artefacts on the seabed (see British Marine Aggregate Producers Association and English Heritage, 2003). Recommendation 29. A number of cultural heritage sites recorded during the field surveys are likely to contain buried human skeletal remains (graves). Inside the site security fence area, such sites are particularly likely to occur within the larger, denser archaeological sites, especially sites in SC5 (and possibly those within SC1 and SC3), which we interpret to be an ancient village site complex (human burials will also almost certainly be present in sites of SC4 and possibly other nearby sites at Konekaru). Archaeological evidence in other parts of the Port Moresby area (e.g. see section 4.5.8.3, section 4.5.8.7), ethnographic records (see section 4.5.7.1, especially section 4.5.7.8, oral traditions and present-day practices indicate that human burials in the recent and distant past were often created beneath or adjacent to house structures, either within hamlets and villages, or at smaller residential localities nearby. Given the presence of village sites within and near the site security fence area – and therefore given the presence of large human populations in the past – it is inconceivable that numerous individuals were not buried in the general area of the proposed developments. We recommend a two-fold strategy to avoid or minimize the chance disturbance of burials during construction of the LNG Facilities: B. David, B. Duncan, M. Leavesley 4-663 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project • Archaeological excavations at the large sites likely to contain human burials (sites at SC1, SC3, and especially SC5) should be extensive enough, and strategically structured, to sample various parts of the site where human burials are most likely to occur underground (i.e. the densest parts of site, with small excavation squares spaced-out across the site). If and when human burials are identified, excavations should expand at those localities sufficiently to exhume the skeletal remains. • During construction, a qualified cultural heritage officer and Boera, Papa, Lea Lea and Porebada village representatives (as appropriate) should on a continued basis monitor earthworks (i.e. ground disturbance) as they are being undertaken for human skeletal remains within the site security fence area and associated infrastructure areas. If human remains are found, a rapidresponse team consisting of a physical anthropologist and appropriate community representatives from Boera, Papa, Lea Lea and Porebada should be at-hand, or ready at short notice (i.e. within 24 hours or less), to salvage such remains to enable the construction activities to continue shortly there-after. It is strongly recommended that strict, formal protocols for the salvaging and management of such human remains, including their extraction from the ground, field and laboratory documentation, community communications, and reburial or museum storage, be discussed between the PNG National Museum and Art Gallery (as administrators of the nation’s major cultural heritage legislation) and Boera, Papa, Lea Lea and Porebada village representatives (representing the contemporary descendants of the region’s ancestral peoples), and subsequently unambiguously written up into a protocol document. This document should then be given to the contractors (and/or their representative delegate organisations) for procedural adoption. Such a document should be made available to the developers at least one month before any ground disturbance associated with the project, to enable all protocol procedures to be set in place prior to salvaging work and prior to any construction. We further recommend that the contractors formally commit to these protocols concerning human skeletal remains, as an explicit commitment of faith to the people of Boera, Papa, Lea Lea and Porebada. Recommendation 30. The indigenous cultural heritage sites within and surrounding the proposed LNG Facilities developments represent a rich array of ancestral places, including ancient and more recent hamlets and villages, story sites, gardens, freshwater wells, clay sources and so forth relating to Koita and Motu history. As noted elsewhere in this report, these are precisely the peoples whose cultural heritage sites, along with those of the Koiari slightly to the east, have already incurred most disturbance as a result of the establishment and continued expansion of the nation’s capital, Port Moresby. While the most significant cultural heritage sites of oral tradition for the Koita and Motu are located not inside the proposed LNG Facilities site development area but rather all around it, the development area none-the-less itself contains a rich assemblage of archaeological sites relating to their history. As a measure of give-and-take and good-will, we recommend a co-ordinated approach to site protection and management, one that would see the establishment of a “protected zone” outside but near the proposed site security fence area, and by which the broader area’s cultural heritage sites and objects can be managed from further industrial developments. While the exact location of this “protected zone” should be defined through discussions between community leaders (including senior clan and section representatives) from each of the four villages (Lea Lea, Papa, Boera, Porebada), PNG National Museum and Art Gallery and National Cultural Commission staff, we suggest here that an appropriate location may be the very highly significant complex of oral tradition and archaeological sites spanning the area from Boera Head in the south to the northern end of Aemakara in the north. There are a number of advantages for making this area a “protected zone”, including: 1) it includes both Koita and Motu ancestral sites; 2) it contains a rich assemblage of sites relating to the hiri, and indeed include many of the most important terrestrial sites relating to the Edai Siabo story about the origins of the hiri; 3) the area is contiguous (Boera) or close (Lea Lea, Papa, Porebada) to existing villages, and therefore as far as logistics are concerned can realistically be managed by local people on a daily basis; 4) the area is today part of a living landscape, with local people using it in everyday B. David, B. Duncan, M. Leavesley 4-664 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project community life, and such community involvements with this landscape should continue to unfold (but not at the expense of LNG Facilities or other large-scale industrial developments within it). The creation of a “protected zone” should not be at the expense of protecting other important oral tradition and archaeological sites elsewhere in the broader region, but would demonstrate good-will and a commitment to looking after significant Koita and Motu historical places while at the same time proceeding with the proposed LNG Facilities developments. Critically, the appropriate village clans/sections must retain management and decision-making control of such a “protected zone”. We recommend that the project commit to the local people to such an initiative, and commence discussions through community, PNG National Museum and Art Gallery and National Cultural Commission round-table negotiations. Such an initiative could take a number of alternative forms, such as a Hiri Centre, or a Koita-Motu historical “protected zone”, and incorporate a keeping place, museum or interpretative centre to hold and display the cultural materials archaeologically excavated during the salvage programme resulting from the LNG Facilities developments. As part of this initiative, we recommend that the project fund the training of “protected zone” personnel to monitor the cultural heritage sites immediately surrounding the LNG Facilities, to ensure their appropriate management in the face of probable long-term indirect and cumulative impacts from the project. We recommend that these personnel continuously monitor and see to fruition the operation of the management plan outlined in Recommendation 4 above. 4.5.14.5 WW11 Sites Recommendation 31. The site of the pipe bridge structure (CB3) between Boera and Konekaru beach should be further investigated to determine its exact location and to record its details. Given that this site was being used during WWII to offload munitions, there is a possibility that UXOs may still exist in this area. We recommend that this site be relocated and recorded prior to construction, with due consideration given to the possible presence of UXOs. Recommendation 32. The P-39 aircraft crash site near Boera (CB4) lies to the south and close to the site security fence area on the edge of the mangrove swamp. An inspection of the site by a cultural heritage specialist in WWII wrecks is recommended to record its exact location, spatial extent and contents. We suggest that any disturbance to this site should only be undertaken in consultation with the DoMH, and probably the Australian and US governments (we recommend specialist advice from the DoMH). Recommendation 33. The proposed shipping channel may affect at least one aircraft wreck (CB23 and/or CB24 in the case of the southern option) known to have occurred in this area. Historical documentary records indicate that other aircraft may have crashed or disappeared in this region, and other previously unreported sites may also exist in the area. A side-scan sonar survey for sunken plane and shipwrecks should be undertaken of the proposed shipping channel route and swing basin area prior to the commencement of any works. Similarly, a side-scan sonar survey should also be undertaken of the Marine Offshore Facility including the associated jetty area if these do not entirely overlap with the offshore areas already surveyed for cultural heritage sites. Recommendation 34. The disturbance of any aircraft crash or shipwreck site should be avoided, particularly if there is the possibility that human remains still exist on-site. The discovery of any WWII ship or plane wreck should be immediately reported to the DoMH, PNG National Museum and Art Gallery, Office of Australian War Graves, and the Joint POW/MIA Accounting Command (JPAC). The discovery of any pre-contact shipwreck site should be reported to the Department of Prehistory, PNG National Museum and Art Gallery. The discovery of any commercial or recreational shipwreck should be reported to the State representatives for Customs, Quarantine and Shipping. Recommendation 35. Any works which would disturb the aircraft crash site at Lea Lea (CB21) should be avoided. An inspection of this site is required by a qualified maritime archaeologist to record the site and document its exact location. B. David, B. Duncan, M. Leavesley 4-665 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Recommendation 36. It should be noted that the U.S. Defence Department has a co-ordinated program to locate, record and exhume the remains of former U.S. servicemen for return to their families. The Joint POW/MIA Accounting Command (JPAC) run world-wide operations with dedicated field teams of forensic scientists and archaeologists to achieve this aim. It has in-country detachments in Vietnam, Thailand and Laos. Similarly, the Office of Australian War Graves (OAWG) also maintains a Register of War Dead and has an undertaking to maintain and commemorate war graves in perpetuity. It is recommended that JPAC and OAWG should be consulted wherever any allied aircraft, shipwreck or other war grave and crash site is discovered, and should also be notified of the aircraft wrecks identified in this report. Recommendation 37. As the planned route for the pipeline is unknown to us, it is possible that other aircraft crash and shipwreck sites at Hidiha (CB22), Bavo (CB28, CB29), and other reefs inside Caution Bay (CB 25, CB26, CB27) may be affected directly or indirectly through construction of the pipeline, and by dredging or localised scouring of the seabed. Many of these sites are considered war graves as some crews died at the crash site and have not been recovered. These wrecks may also include shipwrecks from WWII (with or without UXO that would present a danger to the installation of the gas pipeline). It is therefore recommended that side-scan sonar or multi-beam surveys be undertaken of the proposed pipeline route to survey/interpret the presence of shipwrecks along the proposed route. For deeper offshore sections, it is possible that this data already exists as part of the seabed surveys associated with off-shore prospecting. If this is the case, then these surveys should be interogated by a maritime archaeologist to inspect for the presence of wrecks. The shallower inshore waters of Caution Bay and other sections of the pipeline route should be inspected using a side-scan sonar survey, and any anomalies should be inspected by commercial divers and a maritime archaeologist prior to the commencement of any works. If any aircraft remains are discovered, works should cease immediately and the site should be reported to the DoMH. Recommendation 38. The Boera battery (CB30) to the north of Boera village should not be subject to any developments (including infrastructure developments such as access roads and the like). No disturbance should take place at this site, as it is considered one of the most intact and undisturbed WWII sites in the country and may also contain remains of pre-colonial period occupation. However, it is very likely that this site will witness increased levels of visitation as a result of the proposed developments. Further, systematic archaeological recording of this site is recommended to determine its historical and pre-colonial period indigenous archaeological significance. Recommendation 39. The Boera battery (CB30) is known to have used the sea areas of Caution Bay and the adjacent LNG152 area for heavy artillery target practice. It is therefore highly probable that 155 mm artillery projectile heads and other rounds from small munitions and ack-ack (anti-aircraft) weapons will be found in this area. Furthermore, the presence of the battery was also likely to have drawn enemy attention, as evidenced by 500 lb bombs found close to the battery. The location of a sea mine-field close to Haidana Island (at the southern end of Caution Bay) also raises the possibility of the presence of rogue sea mines. It is highly probable that other UXO such as artillery projectiles, bullets and bombs may be found in both the underwater and terrestrial sections of the LNG152 area, some of which may be buried below ground or the seabed. It is therefore recommended that a UXO survey be carried out at these areas by qualified personnel before the commencement of any construction works in the area. Recommendation 40. The Australian War Memorial is likely to hold extensive further information regarding war-time activities in the Caution Bay area. Further investigation of this collection is recommended to ascertain what records still survive relating to the PNG WWII campaign (and associated military cultural heritage sites) that may exist in this and other areas that may be affected by the installation of the proposed pipeline. Furthermore, the map and chart collections/pilots’ records at the National Library in Canberra should also be searched to ascertain the existence of any plotted B. David, B. Duncan, M. Leavesley 4-666 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project maritime historical features in the area of the proposed LNG152 developments and their associated infrastructure, including the pipeline. Recommendation 41. We recommend that further historical research be undertaken at the Australian War Memorial to determine the exact location and route of the underwater communications cable between Australia and PNG (CB5), and whether the cable is still operational/functional as a telecommunications line. The location where the cable came ashore should be located and recorded, and its location marked with a suitable heritage plaque. 4.5.14.6 Communications Recommendation 42. A major concern that was raised by Lea Lea, Papa, Boera and Porebada villagers at the time of the cultural heritage site surveys is that communication with local communities needs to be improved. We recommend that regular discussions be held with community representatives and community representative groups concerning cultural heritage matters. This includes communicating promptly to the local communities the cultural heritage issues and recommendations made in this report, through 1) a public presentation and discussions via an independent cultural heritage officer (not a co-venture participant worker) accompanied by a coventure participant community liaison officer (who could report directly back to the co-venture participants) within each of the four villages; 2) a written community report; and 3) a permanent or semi-permanent poster stand at each village. We also recommend that a temporary independent cultural heritage officer be employed by which two-way communication can be undertaken between the communities and the developers on an on-going basis. This officer would need to have regular access to a vehicle to enable efficient communications with villagers. This position should be replaced by the cultural heritage officer of a “protected zone” discussed in Recommendation 30 above once this position becomes established. Recommendation 43. We recommend that an independent, trained cultural heritage officer be employed (see also Recommendation 42 above), initially in Port Moresby but subsequently based at the keeping place or museum of a “protected zone” mentioned in Recommendation 30 above, by which villagers can communicate issues concerning cultural heritage matters to the developers, and by which the developers can discuss issues to community groups. It is important that this individual be independent of the development so as to foster neutral ground from which discussions relating to cultural heritage matters can be broached from all parties (e.g. training needs, management concerns, tourism opportunities, potential controlled access to specified sites for LNG Facilities workers etc.). 4.5.14.7 Conclusion Subject to the above recommendations being met, and subject to the completion of the cultural heritage sites surveys for the area within the southeast section of the site security fence area (where surveys have not yet been undertaken, and therefore we cannot make statements about this section’s cultural heritage significance) being free of ‘show stoppers’, and as far as cultural heritage sites are concerned, we recommend that the proposed LNG Facilities and associated shallow marine infrastructure developments as specified in this report can proceed without further cultural heritage sites constraints. B. David, B. Duncan, M. Leavesley 4-667 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project ANNEXURE 4.5 Portion 152 Land Cultural Site Recording Form 4.5 Annexure – Portion 152 Land Cultural Site Recording Form 4-668 Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Site #: ________ PNG National Museum Site #: ________ Site name: __________________________ Other site name(s): __________________ ( ____________ ) Site owner(s): ______________________ Language area: _____________ Clan: ____________________________________________________ Recorded by: _______________ People present: ________________________ Date: ____ / Dec. 2008 GPS location – Grid reference: ____________________ ___________________  AGD66  WGS84 Latitude: _________________ S Longitude: _________________E GPS make: ________________ Elevation ASL: _______ m Locational accuracy: ________ m Vine-Thicket Grassland Open Woodland Forest Rainforest Mangrove Garden Other Surface sediment type at site:  Clay  Silt  Sand  Loam  Gravel  Rock  Freshwater  Saltwater  Other:____________ Ground surface visibility at site:  0%  1-25%  25-50%  50-75%  75-100% Bedrock:  Exposed  Limestone _____________ 4.5 Annexure – Portion 152 Land Cultural Site Recording Form  Not exposed  Other: Other Beach River Levee Floodplain Plain water Swamp Swamp Mangrove/Salt Freshwater River/Creek Spring Lake/Waterhole Lakeside Sinkhole Rocky Outcrop Cliff Base of Hill Hill Slope Hill Top Site Vegetation Ecotone Site Landscape Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Site type: Cultural materials present: Stratification:  Open Shell Midden  Ochre  Surface Site  Rock-shelter  Hearth/earth oven  Stratified/Buried Site  Cave  Charcoal/Ash  Site Exposed by  ‘European’/Asian Contact Materials  Unknown Disturbance  Stone Arrangement  Shell Arrangement  Bottle Glass  Isolated Shell  Metal Items  Grinding Stone  Ceramic  Isolated Stone/Ceramic  Other: ___________  Cultural Materials Scatter  Stone Artefacts  Quarry  Quartz  1-10  11-50  50+  Rock-art  Volcanic  1-10  11-50  50+  Burial/Ossuary  Metamorphic  1-10  11-50  50+  Garden  Sedimentary  1-10  11-50  50+  Temporary Encampment  Shell MNI:  1-10  11-50  50+  Village  Melanesian Ceramics  1-10  11-50  50+  Mound  Animal Bone  1-10  11-50  50+  Ritual/Spirit/’Sacred’ Site  Human Bone  Fishing/Hunting Site  Wooden Structures/Posts/Post-Holes  Sago Processing Site  Macrobotanical remains: _____________________________________  Tree  Culturally Altered  Other: ____________________________________________________  Other: ______________________________________________________________________________ Human skeletal remains: Rock-art: Skulls MNI:  1-10  11-50  50+  Paintings  1-10  MNI:  1-10  11-50  50+  Drawings  1-10  11+ Mandibles 11+ 4.5 Annexure – Portion 152 Land Cultural Site Recording Form Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Postcranials MNI:  1-10  11-50  50+  Stencils  1-10   Prints  1-10   Engravings  1-10   Other  1-10  11+ 11+ Shells:  Anadara sp.  Polymesoda sp.  Syrinx sp. 11+  Cypraea sp.  Oliva sp.  Conus sp.  Melo sp. 11+  Other Bivalve ___________  Other Gastropod _________ Modified Shells: _____________________________________________  1  2-5  5-10  11+ Distance to nearest freshwater source: Vertebrate remains Closest Permanent Water: __________ m  Bird  Swamp  River  Creek  Waterhole/Lake  Other: _______ Closest Temporary Water: __________ m  Swamp  Creek  Waterhole/Lake  Other: _______ Distance to High Water Mark: __________ m 2 Site area: _______ m  Fish  Turtle  Dugong  Pig  Other Mammal Site length (max): ______ m Site Width (max): ______ m  Other: _____________ Site height (max) (for rockshelters/caves): ______ m Related information recorded:  Photos: How many? _____  Drawings: What? ____________________________________  Audio: Who? _______________________________________ Notes: _________________________________________________________________________________ _________________________________________________________________________________ ________ 4.5 Annexure – Portion 152 Land Cultural Site Recording Form Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Environmental description (include information on vegetation assemblage, landforms present, slope, topographic irregularities, drainage): _________________________________________________________________________________ _________________________________________________________________________________ _________________________________________________________________________________ _________________________________________________________________________________ ________________ Site description: ______________________________________________________________________ _________________________________________________________________________________ _________________________________________________________________________________ ________ _________________________________________________________________________________ _________________________________________________________________________________ ________ Management Recommendations: 2 2 2  Large Site (>1000m )  Medium Site (26-1000m )  Small Site (<26m )  Protect  Further Mapping Required  Other:  Salvage __________________________ Comments: ___________________________________________________________________________ _________________________________________________________________________________ ____ 4.5 Annexure – Portion 152 Land Cultural Site Recording Form Social Impact Assessment 2008 Papua New Guinea Liquefied Natural Gas Project Worthing, M.A. 1980. South Papuan coastal sources of potsherds from the Gulf area of PNG. Oral History 8(8):87-100. Wurm, S. & Hattori, S. (1981) Language Atlas of the Pacific Area. Australian Academy of the Humanities and the Japan Academy: Canberra. Dr L R Goldman 7-23