Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Tim Denham,1 Simon Haberle2 & Carol Lentfer3

This review of the evidence for early agriculture in New Guinea supported by new data from Kuk
Swamp demonstrates that cultivation had begun there by at least 6950-6440 cal BP and probably
much earlier. Contrary to previous ideas, the first farming in New Guinea was not owed to SouthEast Asia, but emerged independently in the Highlands. Indeed plants such as the banana were
probably first domesticated in New Guinea and later diffused into the Asian continent.
Keywords: Mid-Holocene, New Guinea Highlands, agriculture, Musa bananas

Introduction
Several areas in the Highlands of New Guinea, in particular the Upper Wahgi Valley, are
significant for interpreting the emergence of agriculture during the early to mid-Holocene.
Of these sites, Kuk Swamp (hereafter referred to as Kuk) is the most important because it
has been investigated in greatest detail, provides the longest chronology and is the type-site
for the investigation of prehistoric plant exploitation in the interior of New Guinea (Denham
2003a; Denham et al. 2003; Golson 1977a, 1982, 1985, 1990, 1991a, 1991b; Golson &
Hughes 1980; Hope & Golson 1995). The international significance of Kuk derives from
the presence of successive phases of drainage for agriculture dating from 6950-6440 cal BP
or earlier (Golson 1977a, 1991a, 1991b; Golson & Hughes 1980; cf. Denham 2004; Denham
et al. 2003; Table 1, see http://antiquity.ac.uk/ProjGall/denham). The evidence at Kuk has
provided a basis for the interpretation of early and independent agricultural origins in the
Highlands of New Guinea.
In this paper, new multi-disciplinary findings on early and mid-Holocene plant exploitation
at Kuk, including archaeological, archaeobotanical and palaeoecological research, are presented
and interpreted with respect to the prehistory of the Upper Wahgi Valley. These results endorse
previous interpretations that New Guinea was a centre of independent agricultural origins,
but question previous ideas that agriculture was necessarily developed in the lowlands and
brought into the Highlands with climatic amelioration and stabilisation at the beginning of
the Holocene (Hope & Golson 1995: 818-819, 827-8). According to our alternative argument,
1 Department of Archaeology, School of Humanities, Flinders University, GPO Box 2100, Adelaide SA 5001, Australia.
(Email: Tim.Denham@flinders.edu.au)
2 Resource Management in Asia-Pacific Program, Research School of Pacific and Asian Studies, Australian National
University, Canberra ACT 0200, Australia. (Email: Simon.Haberle@anu.edu.au)
3 Centre for Geoarchaeology and Palaeoenvironmental Research, School of Environmental Science and Management,
Southern Cross University, PO Box 157, Lismore NSW 2480, Australia (Email: clentfer@scu.edu.au)
Received: 25 November 2003; Revised: 4 May 2004; Accepted: 9 May 2004

839

Research

New evidence and revised interpretations


of early agriculture in Highland New
Guinea

New evidence and revised interpretations of early agriculture in Highland New Guinea

agricultural practices arose in the Highlands of New Guinea during the early to mid-Holocene
through the transformation of pre-existing plant exploitation strategies (following Denham
& Barton in press; Haberle 1993: 299-305); the nature and timing of the earliest innovations
require clarification through further research.
The early and mid-Holocene remains, corresponding to Phases 1 and 2 at Kuk, are directly
relevant to understanding the origins of agriculture in New Guinea. The early and mid-Holocene
remains pre-date any known Southeast Asian influence on New Guinea after its initial colonisation
by at least c. 40 000 cal BP (Groube et al. 1986). Mid-Holocene Southeast Asian influences are
usually associated with the expansion of Austronesian language-speakers to Melanesia at 35003300 cal BP (see Bellwood 1997: 210-54), or at c. 3200-3300 cal BP (Spriggs 2001: 240). The
influence of Austronesian diffusion on mainland New Guinea, marked by the Lapita cultural
complex in island Melanesia (Green 1991), is unclear because there is a near absence of Lapita
pottery and other well-dated traits on the island of New Guinea (see Terrell & Welsch 1997 for
the exceptions). On linguistic grounds, Austronesian influence on New Guinea is considered to
post-date the onset of Lapita culture by at least 1000 years (Bellwood 1996: 487).

Site location
The wetland archaeological site at Kuk Swamp is located approximately 12 kilometres (km)
north-east of the town of Mount Hagen in the Upper Wahgi Valley, Western Highlands
Province, Papua New Guinea (Figure 1). Kuk is located in the Highlands, i.e., land above
1200 metres above mean sea level (m), at an altitude of c.1560 m. Kuk Swamp forms part of
extensive wetlands carpeting the floor of the Upper Wahgi Valley, which is part of one of the
largest inter-montane valleys in the interior of New Guinea. The archaeological site is located
on the former Kuk Agricultural (originally Tea) Research Station that covered 280 hectares
(ha). Approximately 100 ha of the southeast corner were subject to multi-disciplinary
investigation (Golson 1977a: 610).
Like other wetlands in the Highlands of Papua New Guinea, Kuk was artificially drained
in the late 1960s and early 1970s for planting. Prior to drainage some parts of the wetland
were underwater and surface gradients were generally less than 1 in 200 (Golson 1977a:
612), except for scattered low mounds of older, tephra-mantled lahar deposits over and around
which the wetland had accumulated (Pain et al. 1987: 268).
The Upper Wahgi Valley has a lower montane humid climate with an average annual
temperature of 19C and annual rainfall of c. 2700 mm (Hughes et al. 1991: 229). Climate
in the Upper Wahgi Valley is dominated by local orographic effects (Powell et al. 1975: 2). A
slight dry season occurs in the Upper Wahgi Valley between May and June, although soil
water content does not usually limit plant growth (McAlpine et al. 1983: 74, 137).

Problems with previous research


Previous claims and evidence for early and independent origins of agriculture in New Guinea
have been treated cautiously (Harris 1996: 568-9; Richerson et al. 2001: 400; Smith 1998:
142-3), or been ignored (Harlan 1995). Doubts have been fostered by problems with the
limited archaeological, archaeobotanical and palaeoecological evidence presented. Only limited
archaeological evidence associated with Phases 1 and 2 at Kuk was previously published (Bayliss840

Research

Tim Denham, Simon Haberle & Carol Lentfer

Figure 1 Map of New Guinea showing the location of important archaeological and palaeoecological sites and those mentioned
in the text. The inset depicts wetland sites in the Highlands (amended with permission from Mountain 1991: Figure 2.1).

Smith 1996: 508; exceptions being Golson 1976, 1977a: 613-4, 1991a, 2000; Golson &
Hughes 1980) and interpretations of the Phase 1 evidence changed through time (BaylissSmith 1996: 508-9). Consequently, the antiquity, mode of formation and function of
palaeochannels and constituent features on the Phase 1 and 2 palaeosurfaces were uncertain
(after Smith 1998: 142-3 and Spriggs 1996: 528). Findings contemporary with Phase 2 were
reported at other sites in the interior, with some published in detail, e.g. Mugumamp (Harris
& Hughes 1978), although all were interpreted in terms of the Kuk sequence (Denham 2003b).
Furthermore, the extent of landscape change coincident with the early and mid-Holocene
drainage phases at Kuk and anticipated to accompany agro-ecosystems was unknown (Spriggs
1996: 528). Geomorphological investigations indicated accelerated rates of deposition on
the wetland margin from the early Holocene, which were suggestive of erosion following
vegetation clearance for swidden-type dryland cultivation in the catchment (Golson & Hughes
841

New evidence and revised interpretations of early agriculture in Highland New Guinea

1980: 296-7; Hughes et al. 1991: 233-234). Powells pollen diagrams from three other sites
in the Upper Wahgi Valley, at Draepi-Minjigina, Lake Ambra and Warrawau, indicated that
forest clearance had occurred on a regional scale by the mid-Holocene (Powell 1982b: 218);
the commencement and intensity of these activities were unknown because there was an
absence of a terminal Pleistocene and early Holocene component in the relevant Upper
Wahgi Valley pollen diagrams. Deposition rates at Kuk, as well as Haberle et al.s (1991)
evidence of cumulative forest clearance pre-dating 7800 cal BP in the Baliem Valley, suggested
the Wahgi clearances could plausibly date to the early Holocene and be synchronous with
Phase 1 archaeological remains at Kuk (Hope & Golson 1995: 823-5).
In addition, there was a near-absence of published archaeobotanical evidence of former crop
plants, domesticated or otherwise (Bayliss-Smith 1996: 508; Spriggs 1996: 528). Macrobotanical
remains of several edible plants and Musa banana phytoliths were documented for archaeological
phases at Kuk (Powell 1982a; Wilson 1985). The associations of these botanical remains were
uncertain because the plants could have dispersed independently of intentional human practices.
Each of these problems has been redressed by the recent investigations at Kuk.

Archaeological evidence of cultivation practices


Over 200 trenches were excavated at Kuk to delimit field systems and house sites (see Figure 2).
Most excavations investigated later ditch networks (Phases 3-6) and comparatively few explicitly

Figure 2 Plan of the south-east of Kuk Station showing the extent of excavations directed by Jack Golson in 1972-7 and by
Tim Denham and Jack Golson in 1998-9.

842

targeted early to mid-Holocene remains (Phases 1-2). The excavations undertaken in 1998 and
1999 were intended to clarify contentious archaeological finds from previous work and to
obtain samples for multi-disciplinary analyses.

Phase 1 at 10 220-9910 cal BP


Archaeological evidence unique to Kuk was claimed to represent management of the wetland
margin for agriculture at c.10 000 cal BP (Golson 1977a: 613-5; 1989: 679; 1991a; Golson
& Hughes 1980; Hope & Golson 1995: 824). Formerly, the archaeology of Phase 1 consisted
of chronologically and perhaps functionally associated (Golson 1981: 56) palaeochannel and
palaeosurface. The palaeochannel was 2 m wide and 1 m deep and traced for at least 500 m
across the wetland (Golson 1977a: 613-5; 2000: 232). A human origin for the palaeochannel
was inferred from several characteristics including directness of line between points of abrupt
change in direction, one of them in a middle of a low hillock of volcanic ash (Golson 1991a:
485) and by analogy with later palaeochannels, which were similarly interpreted to be artificial
(Golson 2000: 232). More recent research has challenged this interpretation, and here the
early palaeochannel is not considered to be artificial (Denham 2004: 47-53). The aggregates
and abundant macrobotanical remains, as well as pollen and microcharcoal signals, within
palaeochannel fills are consistent with forest disturbance using fire and erosion within the
catchment (see below).
The palaeosurface, by contrast, consists of pits, runnels, stake- and post-holes associated with
sparse artefacts, relatively heterogenous feature fills and an immature soil profile (Figure 3b).
The artificiality of palaeosurface features has been determined from the morphology of features
which have well-defined, cut edges. The inferred functions of features are consistent with planting
and digging (pits), localised overland flow (runnels) and staking of plants (stake and post holes)
within a cleared plot (after Powell et al. 1975: 24-6). The slightly heterogeneous feature fills
and pedogenesis are consistent with limited mechanical admixture and the formation of immature
A/C profiles within an alluvial setting (after Collins & Lavney 1983; Haywood 1985). Lastly
the artificial character of the palaeosurface is confirmed by the presence of stone tools, which
display evidence for the processing of starchy and woody plants and include a flake used to
process taro (Colocasia esculenta) (Denham et al. 2003: 191).
Based on excavations to date, features are restricted to higher ground adjacent to the
palaeochannel. This area was only dry enough to utilise for a short period, perhaps only
once, prior to abandonment. In contrast to previous portrayals, these features need not
represent specialised wetland management, but may be the spatial extension of plant
exploitation practices from adjacent dryland slopes.
The plant exploitation practices represented by early Holocene features are unclear. Features
could be associated with relatively adventitious plant management and use, although from a
different perspective they could be consistent with practices in an inter-cropped, swiddentype plot on the wetland edge. Without additional investigations, it is not possible to determine
if these features represent agriculture, as previously claimed, or other activities.

843

Research

Tim Denham, Simon Haberle & Carol Lentfer

Figure 3 Archaeostratigraphic summary of Phases 1 and 2 at Kuk (modified with permission from Denham et al. 2003: Figure 2). Field plans were drawn by Arthur and Cherie Rohn.
Photographs were taken by E.C. Harris in 1977 and reproduced courtesy of Jack Golson, Records of the Australian Museum (photo b) and Archaeology in Oceania (photo c).

New evidence and revised interpretations of early agriculture in Highland New Guinea

844

Tim Denham, Simon Haberle & Carol Lentfer

Mid-Holocene finds at Kuk represent the earliest, unequivocal evidence for agriculture in
New Guinea (Denham et al. 2003; Golson 1977b: 48-9, 1989: 679, 1991a: 484, 489; Hope
& Golson 1995: 823). Phase 2 at Kuk was formerly characterised as consisting of at least
four successive palaeochannels that articulated with palaeosurfaces comprised of interconnected networks, or more dispersed features (Golson 1977a: 615-9). As with the early
Holocene palaeochannel, there is insufficient evidence to determine that the mid-Holocene
watercourses were human-made (Denham 2003b: 163-4). Palaeochannel characteristics,
including slightly sinuous planforms, cross-sectional morphologies and long-profile gradients,
are arguably consistent with unmodified watercourses.
The Phase 2 palaeosurface has well-connected (integrated) and less-organised (discrete)
areas. The integrated areas have been studied in detail and consist of regular-to-semi-regular
networks of cut features defining upraised areas, previously referred to as raised island beds
(Golson 1977a: 616), against the underlying grey clay (Figure 3c). Similar mid-Holocene
palaeosurfaces are preserved at Warrawau (Golson 2002) and Mugumamp (Harris & Hughes
1978) in the Upper Wahgi Valley, with proximal locales at Kana (Muke & Mandui 2003)
and Ruti (Gillieson et al. 1985) providing more equivocal evidence (see Figure 1; reviewed in
Denham 2003b: 169-74). The preserved island beds represent the truncated bases of former
mounds constructed on the wetland margin for cultivation. At the time of initial formation,
grey clay was probably a B-horizon, with the original topsoil only surviving as displaced
remnants within feature fills. X-radiographs and thin sections depict greater heterogeneity
within fills in comparison to adjacent stratigraphic units, which lends support to previous
interpretations of mound construction.
Artefacts collected from Phase 2 contexts are sparse, which is anticipated at an agricultural
site in Highland New Guinea where wooden tools were more probably used (Golson 1977c).
Lithic artefacts were used to process Colocasia taro, probable Zingiber gingers and other
unidentified plants (Denham et al. 2003: 191-2, Table S3). Additionally, one palaeosurface
feature exhibits anomalously high Musaceae (banana) phytolith frequencies, including seed
phytoliths of Eumusa, which are interpreted to represent deliberate planting within grassland
maintained by burning (Denham et al. 2003: 191-2, Figure 3).
Previous interpretations suggested that mounded cultivation enabled the multi-cropping of
plants with different edaphic requirements (Golson 1977a: 617). Water-tolerant plants, e.g.
Colocasia taro, could potentially have been planted within the damper features between mounds,
whereas water-intolerant plants, e.g. sugarcanes (Saccharum spp.), Musa bananas, yams (Dioscorea
spp.), edible pitpit (Setaria palmifolia) and mixed vegetables could have been planted and staked
on the island beds (Golson 1977a: 616, 1981: 57-8; Powell et al. 1975: 42). Archaeological
and archaeobotanical evidence is consistent with this agricultural interpretation.

Palaeoecology and landscape transformation


Palaeoecological studies in New Guinea rely on a series of criteria to distinguish human
impacts from natural processes in a record of vegetation history. These primarily include the
identification of processes that did not previously occur in the palaeoecological record such
as indications of forest decline and burning, and increases in secondary forest and herbs
845

Research

Phase 2 at 6950-6440 cal BP

New evidence and revised interpretations of early agriculture in Highland New Guinea

(Haberle 1994). Pollen


records from Lake Ambra,
Draepi-Minjigina
and
Warrawau
(Mantons)
Plantation in the Upper
Wahgi Valley document
disturbance to the primary
forest that pre-dated 62005700 cal BP (Powell 1970a:
199, 1980: 19, 1981: 306,
1982b: 218; Figure 4). Lake
Ambra
and
DraepiMinjigina signalled a change
from undisturbed primary
forest in the late Pleistocene
to a disturbed environment
with higher percentages of
woody non-forest species by
the mid-Holocene (Zone III;
Powell 1970b: 165-86, 1981:
306). Similarly, the M1 core
from Warrawau depicted the
establishment of secondary
forest from the beginning of
the pollen diagram at 5500
cal BP (Powell 1970b: 1559; Powell et al. 1975: 43-4,
46-8). Secondary forest was
represented by light
demanding species including
Trema sp., Acalypha sp.,
Macaranga sp. and Dodonaea
sp.
These Wahgi Valley sites
record further declines in
forest cover at different times
over the following 4000 years
(Powell 1982b: 218),
suggesting a mosaic and
time-transgressive pattern of
forest clearance activity
throughout the mid and late
Holocene. Other studies
across the Highlands also

Figure 4 Diagram of stratigraphy, inferred chronology and summary arboreal pollen


with Poaceae (pollen sum for all taxa based on total forest and woody non-forest
taxa) from three swamp sites in the Upper Wahgi Valley (redrawn and amended
with permission from Powell 1982b: 219, Powell 1982b: 221 and Powell 1970b:
Figure 8.7, respectively). Solid lines mark comparable pollen assemblages at the three
sites based on pollen assemblages and correspond approximately to inferred ages given
in the original texts: Zone I (before 21 000 cal BP), Zone II (c. 21 000-6000 cal
BP) and Zone III (c. 6000 cal BP-present). Calibration of 14C ages after 19 200 BP
are based on Stuiver and Reimer (1993) and before 19 200 BP are derived from a
linear conversion following Bard et al. (1990).

846

show evidence of midHolocene forest clearance


(Haberle 1994: Figure 8.2).
Of most significance, human
clearance of primary forest at
Kelela Swamp in the Baliem
Valley occurred prior to
c. 7800 cal BP, with
continued disturbance and
the establishment of
extensive grasslands by 3000
cal BP (Haberle et al. 1991).
The early Holocene gap in
the pollen records from the
Upper Wahgi Valley has been
filled by recent research at
Kuk. Samples collected from
a core extending from the
Late Pleistocene to late
Holocene, as well as those
collected from the fills of
palaeosurface features and
palaeochannels, enable
detailed reconstructions of
palaeoenvironments during
the early and mid-Holocene
(Figure 5). A composite
diagram incorporating a
previously unpublished
pollen diagram from Kuk
(Powell 1984) illustrates the
swamp vegetation history
from before the Last Glacial
Maximum (LGM) through
to the present (Figure 6).
Cold-adapted vegetation
in the form of grass and sedge
dominated
wetlands
persisted through to the early
Holocene at least partially
under the influence of fire
(zone K-1). In the early
Holocene (zone K-2), the

Research

Tim Denham, Simon Haberle & Carol Lentfer

Figure 5 Composite pollen diagram showing selected pollen and spore taxa (per cent
base on total pollen and spore sum excluding aquatics) plotted as sample numbers
and ordered according to archaeostratigraphic associations (left of diagram). The
Musaceae phytolith frequency curve is derived from the per cent of total phytoliths
(counted separately). Significant pollen zones (K-1 to K-4) are shown on the right
side of the diagram. Calibrated age ranges based on Denham et al. 2003: Tables S1
and S2.

847

New evidence and revised interpretations of early agriculture in Highland New Guinea

Figure 6 Palaeoenvironmental reconstruction for Kuk Swamp catchment based on a composite pollen diagram of archaeological
and stratigraphic samples (see Figure 5) and a previously unpublished core in A10 f/g (amended with permission from Powell
1984). The diagram correlates Powells pollen zones at Kuk and other Upper Wahgi Valley sites (I-III) with Haberles pollen
zones (K-1 to K-4) at Kuk.

mixed-montane forest is replaced by a more open forest environment with Pandanus


(probably P. antaresensis) and Schefflera as important swamp forest components. Between 10
200 cal BP and 7400 cal BP (zones K-2 and K-3), and during deposition of a distinctive
grey-brown clay on the wetland margin, grasslands and fern flora increase at the expense of
forest which is under the influence of periodic fire episodes. At the same time, the catchment
forest composition changes from dominance of montane canopy taxa, such as Nothofagus,
Castanopsis and gymnosperms to subcanopy taxa, particularly Pandanus. At the beginning of
the mid-Holocene, at around 7000 cal BP (zone K-4), there is a rapid loss of forest associated
with increased burning and an open grass-sedge swampland is established.
The new palaeoecological evidence from Kuk, in conjunction with previous diagrams at
Draepi-Minjigina, Lake Ambra and Warrawau, indicates that forest clearance, widespread
use of fire and the establishment and maintenance of disturbed environments were regional
processes across the Upper Wahgi Valley from at least 7000-6500 cal BP. Taken in conjunction
with archaeological evidence for prehistoric mounding at Kuk, Mugumamp and Warrawau,
this vegetation history represents the establishment of an agricultural landscape in the Upper
Wahgi Valley by the mid-Holocene.
848

Tim Denham, Simon Haberle & Carol Lentfer

Archaeobotanical investigations at Kuk employed macrofossil (e.g. seed, wood) and microfossil
(e.g. phytolith, pollen, starch grain) techniques. The archaeobotanical work documents a
wide range of traditional food plants in the Kuk vicinity at various times in the past (Table 2,
at http://antiquity.ac.uk/ProjGall/denham). Residue analysis of the used edges of stone tools
from early and mid-Holocene contexts indicates that Colocasia taro, cf. Zingiber gingers, and
other starchy and woody plants were being exploited (Denham et al. 2003: 191).
The most significant early Holocene additions to the suite of edible plants present in the
Kuk vicinity are Colocasia taro and Eumusa bananas (Denham et al. 2003: 191-2; cf. Wilson
1985: Table 3, at http://antiquity.ac.uk/ProjGall/denham). Colocasia taro has been identified
from starch residues on the cutting edge of a flake collected from the fill of an early Holocene
feature, as well as on artefacts collected from within grey clay and the fill of a Phase 2 feature
(Denham et al. 2003: 191). Eumusa section banana phytoliths are present at the base of grey
clay sealing a 10 000 year-old palaeochannel (see Figure 5). Intact Musa phytolith chains
within grey clay contexts are significant because they signify a local origin. Articulated chains
of Musa phytoliths are fragile and are suggestive of local in situ decomposition and reflect
limited post-depositional disturbance and translocation. Peaks within early Phase 2 contexts
are, given their archaeological associations, more important and show banana phytoliths,
including those of Eumusa section, accounting for up to 9 per cent of all phytoliths. Given
the low productivity of phytoliths in bananas relative to grasses, elevated frequencies in Phase
2 contexts have been interpreted to represent the cultivation of bananas on the wetland
margin within a landscape degraded to grassland (Denham et al. 2003: 191-2).
In New Guinea, Colocasia taro is presumed to be a lowland plant due to greater phenotypic
variation among contemporary wild populations at lower altitude (Yen 1991: 77, 1995:
835). The present-day altitudinal range of Colocasia taro is assumed to be a product of
agronomic selection (Yen 1995: 843). Similarly, wild Eumusa banana populations only occur
at lower altitudes than Kuk in New Guinea (Argent 1976; Simmonds 1962), although like
Colocasia taro, they may have spread to higher altitudes independently of human action as
post-LGM climates ameliorated.
Although the mechanisms for the diffusion of Colocasia taro and Eumusa section bananas
to the Highlands are unclear, these plants were potentially important sources of starch for
inhabitants in the early Holocene. Earlier occurrences of these and other plants may yet be
found given the absence of artefacts and the lack of detailed multi-disciplinary investigations
of stratigraphy dating to the Late Pleistocene at Kuk and other Highland sites. The lack of
correlation between archaeobotanical remains and archaeological phases at Kuk indicates
that plant exploitation was not restricted to the wetland margin during the early and midHolocene. Most edible plants in the Highlands are traditionally cultivated asexually from
wild populations in the vicinity and, therefore, many are anticipated to be continually
represented in the archaeobotanical record.
Colocasia taro and Eumusa bananas, along with breadfruit (Artocarpus altilis), sago
(Metroxylon sagu), sugarcanes (Saccharum spp.) and some yam species (Dioscorea spp.), are
thought on phytogeographic and genetic grounds to be indigenous to, and domesticated in,
the New Guinea region (De Langhe & de Maret 1999; Lebot 1999; Matthews 1991, 1995,
2003b). Indeed, Lebot states (1999: 626): New Guinea and some parts of Melanesia have
849

Research

Archaeobotanical evidence of plants

New evidence and revised interpretations of early agriculture in Highland New Guinea

been areas of domestication and still are important centres of diversity for species that also exist in
Asia. The archaeobotanical findings from Kuk and other sites for Colocasia taro (Haberle
1995; Loy et al. 1992) provide strong corroborating evidence for these interpretations and
suggest that the New Guinea region was a major centre of plant domestication.

Reconsidering the emergence of agriculture in New Guinea


Timing: early or mid-Holocene?
Early evidence of occupation at several sites in the Highlands dates to the terminal Pleistocene
and early Holocene (Golson 1991b: 88-9). Although, in part, this trend may be apparent
and reflect a temporal-decay effecy, i.e. comparatively fewer open sites survive with increasing
age, it may also reflect greater populations of mobile groups living in the Highlands. Based
on the inventory at Kuk, a broad range of edible plants was present in the Highlands during
the Late Pleistocene (see pre-P1 in Table 2), which together with hunting and foraging for
insects and grubs could have facilitated permanent occupation of higher and lower altitude
regions of the interior following initial colonisation of the island (cf. Groube 1989).
Throughout this period, mobile groups may have ranged across the interior via the expansive
glacial grasslands that connected inter-montane valleys (Hope & Hope 1976) and utilised
lowland resources, such as sago (Metroxylon sagu), as well as those in the Highlands.
At the end of the LGM and prior to stabilisation in the early Holocene, Highland climates
are considered to have fluctuated considerably and to have been subject to greater interannual
variability than is experienced in the current El Nio-related climate regime (Haberle et al.
2001). As suggested for other parts of the world (Bellwood 1996), fluctuating climates after
the LGM may have stressed existing plant management practices in Highland New Guinea
and necessitated more interventionist (i.e. greater ground preparation, tending and weeding
of plants) and more extensive (i.e. greater disturbance of forest canopy to increase habitat
diversity) plant exploitation strategies in order to maintain yields and a broad-spectrum diet
(Haberle 1993: 305-8). Such strategies may be represented by the asynchronous and noncumulative records of human-induced forest disturbance using fire in the Highlands in the
Late Pleistocene (e.g. Haberle 1998; Hope 1983; Powell 1984). Additional forcing occurred
during the early and mid-Holocene as people sought to obtain food in increasingly degraded
and resource-poor environments on the floors of the larger inter-montane valleys. The
signatures of more extensive and interventionist strategies are apparent during the early and
mid-Holocene at Kuk, as well as at other archaeological and palaeoecological sites at wetlands
in the Highlands.
During the early Holocene (10 200-9900 cal BP) at Kuk, and perhaps elsewhere in the
Upper Wahgi Valley, people used fire to increasingly disturb and modify the mixed montane
forest and soils on the valley walls. The resultant landscape comprised a continually changing
mosaic of primary and secondary forest, regrowth and grasslands on the dryland slopes and
variously disturbed riparian and wetland environments. Several high-caloric and nutritious
plants were present in the Kuk vicinity including Castanopsis spp., Colocasia taro, Musa bananas
(including identifiable members of Eumusa section with the onset of grey clay deposition),
Pandanus spp., cf. Setaria palmifolia, and cf. Zingiberaceae. These plants, supplemented by
hunting and foraging, could have been sufficient to sustain broad-spectrum diets for permanent
850

occupants of the Highlands. At this time, people were adventitiously exploiting slightly raised
and drier ground adjacent to a palaeochannel at Kuk, where there is evidence suggesting
digging and staking within a plot and the exploitation of Colocasia taro and other starchy
and woody plants.
By the mid-Holocene (6950-6440 cal BP), the forests on the floor of the Upper Wahgi
Valley had been altered to grassland or to secondary regrowth. Within relatively nutrient
poor grassland environments, people gradually increased their degree of intervention in the
management of significant food plants, especially high-energy staples. Most plants could still
be collected or transplanted from primary and secondary forest and regrowth habitats in the
valley. At this time, mounds were made within cleared plots along the wetland edge and
planted with bananas and other inter-cropped plants. Colocasia taro, cf. Zingiber gingers and
other starchy and woody plants were processed, presumably for consumption.
Multi-disciplinary lines of evidence provide an impression of plant exploitation on the
wetland margin in the early Holocene. Similar palaeoecological signals, but without
accompanying archaeological remains along the wetland edge, persist for the next 3-4000
years during grey clay deposition. Although suggestive of swidden-type cultivation (Denham
2004), the evidence is currently insufficient to determine whether these early Holocene
practices are comparable to contemporary agricultural practices in the region (cf. Bourke
2001; Brookfield & Hart 1971: 94-124; Powell & Harrison 1982; Powell et al. 1975). By
the mid-Holocene, a clearer picture of agricultural practices emerges and their effects are
clearly documented in the cultivation and use of plants (e.g., bananas and Colocasia taro,
respectively), some of which are considered to have undergone domestication in New Guinea,
and in the dramatic transformation of the environment with the degradation of forest to
grassland. Archaeological remains of prehistoric mound preparation on the wetland edge
provide the foundation for interpretations of the archaeobotanical and palaeoecological
evidence. Taken together, the multi-disciplinary lines of mid-Holocene evidence are
comparable to contemporary agricultural practices and their effects in the Highlands of New
Guinea.

Location: Highlands or lowlands?


Lowland environments and inhabitants were not subject to as severe climate-induced stresses
as the Highlands during the terminal Pleistocene and the Holocene. Lowland inhabitants
had continuous access to a broad range of fauna and nutritious and high-caloric food plants.
Of most significance are a range of indigenous nut-bearing (e.g., Artocarpus altilis, Barringtonia
spp., Canarium spp., Cocos nucifera, some Pandanus spp., Terminalia spp.) and starch-bearing
trees (e.g., Metroxylon sagu), which have restricted lowland distributions (Bourke 1996; Yen
1995: 833-40). Additionally, a broad range of terrestrial, riverine, lagoonal and littoral fauna,
including numerous freshwater fish, reptiles and shellfish, have distributions restricted to the
lowlands. The relative abundance of edible tree resources led to an early focus on arboriculture
in the lowlands, possibly from the terminal Pleistocene (Yen 1996: 41). A focus on managed
stands of trees supplemented by hunting, fishing, foraging and garden horticulture created
diverse practices some of which are not readily classified as agricultural or hunter-gatherer
(Dwyer & Minnegal 1991; Roscoe 2002), but are ambiguous in terms of such traditional
terminology (Terrell 2002; Terrell et al. 2003).
851

Research

Tim Denham, Simon Haberle & Carol Lentfer

New evidence and revised interpretations of early agriculture in Highland New Guinea

The diversity of present and, by extension, past practices blurs any clear distinction between
types of Highland and lowland subsistence. In the Highlands, however, there is a greater reliance
on garden horticulture based on the vegetative propagation of vegetables and root-crops. Garden
horticulture is practiced more intensively and forms the subsistence base in the Highlands,
whereas in many lowland locales it supplements arboriculture and other activities.
Multiple lines of archaeological, archaeobotanical and palaeoecological evidence suggest
that agriculture in New Guinea, manifest as extensive (swidden) or intensive (semi-permanent
and short fallow wetland and raised bed) cultivation, originated in the Highlands. Currently,
however, there is a lack of comparable evidence for lowland subsistence, particularly
arboriculture, to test this proposition (for exceptions see Swadling et al. 1991; Yen 1996).
There is little archaeological and palaeoecological evidence to indicate the expansion of
agriculturalists from the lowlands brought plants and horticultural practices into the Highlands
during the terminal Pleistocene or early Holocene. Although the altitudinal range of Colocasia
taro and some Musa bananas lies dominantly within the lowlands implying a lowland origin
for these plants, the upper altitudinal limits lie within or close to the Highland valleys. The
appearance of these starch-rich plants at Kuk at the beginning of the Holocene may relate to
natural dispersal processes (i.e. seed-dispersed under warming climate conditions) as much
as to human agency (i.e. brought by mobile groups engaged in an extensive form of plant
exploitation based on vegetative propagation).

Conclusion
Recent multi-disciplinary investigations have filled out previous, largely speculative
chronologies of prehistoric plant exploitation (Harris 1995) and confirm previous
interpretations that agriculture arose independently in New Guinea (Table 3). New multidisciplinary investigations at Kuk have advanced previous interpretations of prehistoric plant
exploitation in the Highlands of New Guinea in two significant ways.
First, multi-disciplinary lines of evidence (including archaeological, archaeobotanical,
palaeoecological and stratigraphic) show that agriculture was practiced in the Upper Wahgi
Valley by at least 6950-6440 cal BP and probably much earlier. The evidence for earlier plant
exploitation practices at Kuk is insufficient to be definitive of agriculture, as previously claimed
(Golson 1977a: 613-5, 1991a; Golson & Hughes 1980; Hope & Golson 1995: 824). The
timing of the emergence of agriculture in the Highlands of New Guinea requires clarification
through greater interpretative resolution of the Phase 1 record at Kuk and the excavation of
additional sites with evidence of early Holocene plant exploitation.
Second, agriculture may have emerged in the Highlands of New Guinea as opposed to the
lowlands. Golson suggested that agriculture originated in the lowlands and spread with
expanding populations into the Highlands as climates ameliorated at the beginning of the
Holocene (Golson 1991b: 88-9; Hope & Golson 1995: 827-8). It is argued here that
agriculture emerged from broad-spectrum plant exploitation practices in the Highlands which
had enabled permanent occupation of the interior during the Late Pleistocene. Climatic
(Late Pleistocene) and environmental (early to mid-Holocene) forcing of existing plant
exploitation practices led to the development of more interventionist and extensive strategies
in the Highlands. Although the specific mechanisms remain uncertain, an increasing focus
852

on major sources of starch, including Colocasia taro and Eumusa bananas, were central to the
emergence of agriculture in the Highlands.
As with an earlier debate concerning arboriculture (compare Kirch 1989 to Swadling et al.
1991 and Yen 1996), agriculture in New Guinea was traditionally viewed as being of Southeast
Asian origin (e.g. Sauer 1952). The archaeological, archaeobotanical and palaeoecological
findings at Kuk corroborate phytogeographic and genetic interpretations of independent
plant domestication in Melanesia and demonstrate that agriculture emerged in New Guinea
independently of any South-east Asian influence by at least 6950-6440 cal BP. Indeed, the
early Holocene Eumusa section bananas at Kuk solidify previous interpretations for the early
diffusion of domesticated plants from New Guinea and their subsequent influence on the
development of agriculture in Southeast Asia (De Langhe & de Maret 1999: 378-82) and
Africa (Mbida et al. 2000, 2001). Such findings open up new possibilities for interaction
between mainland Southeast Asia and New Guinea during the early and mid-Holocene (i.e.
prior to Austronesian dispersal into Indo-Malaysia and Melanesia) and require us to rethink
the origins and spread of agriculture in the Pacific, Southeast Asia and beyond.
Acknowledgements
The authors are indebted to Emeritus Professor Jack Golson for his assistance, encouragement and enthusiasm
during every stage of the project. Denham acknowledges scholarships and a fieldwork grant from the School of
Archaeology and Anthropology, Australian National University (ANU) and grants from the Australia and Pacific
Science Foundation (to J. Golson in 1998 and 1999) and the Pacific Islands Development Program (in 1999).
Radiocarbon dates were provided by Centre for Archaeological Research, Australian Nuclear Science and
Technology Organisation and Research School of Earth Sciences, ANU (courtesy of John Chappell and Abaz
Alimanovic). Haberle undertook his research during an Australian Research Council QEII Fellowship. Lentfer
acknowledges a scholarship from Southern Cross University and grants from the Pacific Biological Foundation
(2002) and Australian Museum. The authors are indebted to the contributions of Richard Fullagar, Judith Field,
Michael Therin (residue analysis, all University of Sydney), Barbara Winsborough (diatom analysis, Winsborough
Consulting, Texas) and Nick Porch (insect analysis, Monash University) although their work does not appear in
this paper, they are integral members of the research team. The authors thank Kay Dancey, Rob Patat, Art and
Cherie Rohn and Resource Management in Asia-Pacific Program, ANU for assistance with figures. Golson and
Jocelyn Powell kindly granted permission to publish the A10f/g pollen diagram from Kuk and Powell granted
permission to reproduce and modify pollen diagrams from sites in the Upper Wahgi Valley. Atholl Anderson,
Soren Blau, Jack Golson and two anonymous reviewers are thanked for comments on earlier manuscripts. This
research was made possible with the permission and assistance of the Papua New Guinea National Museum and
Art Gallery, National Research Institute, Western Highlands Provincial Government and the Kawelka at Kuk.
BAYLISS-SMITH, T. & J. GOLSON. 1992a. Wetland
agriculture in New Guinea Highlands prehistory,
in B. Coles (ed.), The wetland revolution in
prehistory: 15-27. Exeter: The Prehistoric Society
and Wetland Archaeological Research Project.

Bibliography
ARGENT, G.C.G. 1976. The wild bananas of Papua
New Guinea, Notes from The Royal Botanic Garden,
Edinburgh 35: 77-114.
BARD, E., B. HAMELIN, R.G. FAIRBANKS & A. ZINDLER.
1990. Calibration of the 14C timescale over the
past 30 000 years using mass spectrometric U-Th
ages from Barbados, Nature 345: 405-410.

1992b. A Colocasian revolution in the New Guinea


Highlands?: insights from Phase 4 at Kuk,
Archaeology in Oceania 27: 1-21.
1999. The Meaning of Ditches: Deconstructing the
Social Landscapes of Drainage in New Guinea,
Kuk, Phase 4, in C. Gosden and J. Hather (eds.):
199-231.

BAYLISS-SMITH, T. 1996. People-plant interactions in the


New Guinea highlands: agricultural hearthland or
horticultural backwater? in D.R. Harris (ed.): 499523.

853

Research

Tim Denham, Simon Haberle & Carol Lentfer

New evidence and revised interpretations of early agriculture in Highland New Guinea
DENHAM, T.P. & C. BALLARD (eds.). 2003. Perspectives
on prehistoric agriculture in New Guinea.
Archaeology in Oceania Special Issue.

BELLWOOD, P. 1996. The origins and spread of


agriculture in the Indo-Pacific region: gradualism
and diffusion or revolution and colonisation? in
D.R. Harris (ed.): 465-98.

DENHAM, T.P. & H. BARTON. In press. The emergence


of agriculture in New Guinea: continuity from preexisting foraging practices, In D.J. Kennett and B.
Winterhalder (eds.), Behavioral ecology and the
transition to agriculture. Washington: Smithsonian
Institution Press (to be published late 2004).

1997. Prehistory of the Indo-Malaysian Archipelago.


Honolulu: University of Hawaii Press.
BOURKE, R.M. 1996. Edible indigenous nuts in Papua
New Guinea, in M.L. Stevens, R.M. Bourke and
B.R. Evans (eds.): 45-55.

DENHAM, T.P., S.G. HABERLE, C. LENTFER, R. FULLAGAR,


J. FIELD, M. THERIN, N. PORCH & B.
WINSBOROUGH. 2003. Origins of agriculture at
Kuk Swamp in the Highlands of New Guinea.
Science 301: 189-193.

2001. Intensification of agricultural systems in Papua


New Guinea, Asia Pacific Viewpoint 42(2-3): 219236.
n.d. Altitudinal limits of 220 economic crop species in
Papua New Guinea. Unpublished manuscript on
file at Research School of Pacific and Asian Studies
(RSPAS), Australian National University (ANU).

DWYER, P.D. & M. MINNEGAL. 1991. Hunting in


lowland, tropical rain forest: Towards a model of
non-agricultural subsistence, Human Ecology 19:
187-212.

BROOKFIELD, H.C. & D. HART. 1971. Melanesia.


London: Methuen.

FRENCH, B.R. 1986. Food plants of Papua New Guinea:


a compendium. Privately published book.

BULMER, S. 1975. Settlement and economy in


prehistoric Papua New Guinea: a review of the
archaeological evidence, Journal de la Socit des
Ocanistes 31(46): 7-75.

GILLIESON, D., P. GORECKI & G. HOPE. 1985.


Prehistoric agricultural systems in a lowland
swamp, Papua New Guinea, Archaeology in
Oceania 20: 32-7.

COLLINS, J.F. & F. LAVNEY. 1983. Micromorphological


changes with advancing pedogenesis in some Irish
alluvial soils, in P. Bullock and C.P. Murphy (eds.),
Soil micromorphology: 297-307. Berkhamstead: AB
Academic.

GOLSON, J. 1976. Archaeology and agricultural history


in the New Guinea Highlands, in G. de G.
Sieveking, I.A. Longworth and K.E. Wilson (eds.),
Problems in economic and social archaeology: 20120. London: Duckworth.

DE LANGHE, E. & P. DE MARET. 1999. Tracking the


banana: its significance in early agriculture, in C.
Gosden and J. Hather (eds.): 377-96.

1977a. No room at the top: agricultural


intensification in the New Guinea Highlands, in J.
Allen, J. Golson and R. Jones (eds.), Sunda and
Sahul: prehistoric studies in southeast Asia, Melanesia
and Australia: 601-38. London: Academic Press.

DENHAM, T.P. 2003a. The Kuk morass: multidisciplinary investigations of early to mid
Holocene plant exploitation at Kuk Swamp, Wahgi
Valley, Papua New Guinea. Unpublished PhD
thesis, Australian National University.

1977b. The making of the New Guinea Highlands, in


J.H. Winslow (ed.), The Melanesian environment:
45-56. Canberra: Australian National University
Press.

2003b. Archaeological evidence for mid-Holocene


agriculture in the interior of Papua New Guinea: a
critical review, in T.P. Denham and C. Ballard
(eds.): 159-76.

1977c. Simple tools and complex technology:


Agriculture and agricultural implements in the
New Guinea Highlands, in R.V.S. Wright (ed.),
Stone tools as cultural markers: change, evolution and
complexity: 154-61. New Jersey: Humanities Press
Inc.

2004. Early agriculture in the Highlands of New


Guinea: an assessment of Phase 1 (c. 9000 bp) at
Kuk Swamp, in V.J. Attenbrow and R. Fullagar
(eds.), A Pacific Odyssey: archaeology and
anthropology in the Western Pacific. Papers in honour
of Jim Specht: 47-57. Records of the Australian
Museum Supplement 29, Sydney: Australian
Museum.

1981. New Guinea agricultural history: a case study,


in D. Denoon and C. Snowden (eds.), A time to
plant and a time to uproot: 55-64. Port Moresby:
Institute of Papua New Guinea Studies.

In press. Agricultural origins and the emergence of


rectilinear ditch networks in the Highlands of New
Guinea, in A. Pawley, R. Attenbrough, J. Golson
and R. Hide (eds.), Papuan pasts: Studies in the
cultural, linguistic and biological history of the
Papuan-speaking peoples. Adelaide: Crawford House
(to be published late 2004).

1982. The Ipomoean revolution revisited: society and


sweet potato in the upper Wahgi Valley, in A.
Strathern (ed.), Inequality in New Guinea Highland
societies: 109-36. Cambridge: Cambridge
University Press.

854

1994. Anthropogenic indicators in pollen diagrams:


problems and prospects for late Quaternary
palynology in New Guinea, in J. G. Hather, ed.,
Tropical archaeobotany: applications and new
developments: 172-201. London: Routledge.

1985.Agricultural origins in Southeast Asia: a view


from the east, in N. Misra and P. Bellwood (eds.),
Recent advances in Indo-Pacific prehistory: 307-14.
Leiden: E. J. Bill.
1989. The origins and development of New Guinea
agriculture, in D.R. Harris and G.C. Hillman
(eds.): 678-87.

1995. Identification of cultivated Pandanus and


Colocasia in pollen records and the implications for
the study of early agriculture in New Guinea,
Vegetation History and Archaeobotany 4: 195-210.

1990. Kuk and the development of agriculture in New


Guinea: retrospection and introspection, in D.E.
Yen and J.M.J. Mummery (eds.), Pacific Production
Systems: Approaches to Economic Prehistory: 139-47.
Canberra: Australian National University.

1998. Late Quaternary change in the Tari Basin,


Papua New Guinea, Palaeogeography,
Palaeoclimatology, Palaeoecology 137: 1-24.
HABERLE, S.G., G.S. HOPE & Y. DE FRETES. 1991.
Environmental change in the Baliem Valley,
Montane Irian Jaya, Republic of Indonesia, Journal
of Biogeography 18: 25-40.

1991a. Bulmer Phase II: early agriculture in the New


Guinea Highlands, in A. Pawley (ed.), Man and a
half: essays in Pacific anthropology and ethnobiology
in honour of Ralph Bulmer: 484-91. Auckland: The
Polynesian Society.

HABERLE, S.G., G.S. HOPE, & S. VAN DER KAARS. 2001.


Biomass burning in Indonesia and Papua New
Guinea: natural and human induced fire events in
the fossil record, Palaeogeography, Palaeoclimatology,
Palaeoecology 171: 259-268.

1991b. The New Guinea Highlands on the eve of


agriculture, Bulletin of the Indo-Pacific Prehistory
Association 11: 82-91.
2000. A stone bowl fragment from the Early Middle
Holocene of the Upper Wahgi Valley, Western
Highlands Province, Papua New Guinea, in A.
Anderson and T. Murray (eds.), Australian
Archaeologist: collected papers in honour of Jim Allen:
231-48. Canberra: Coombs Academic Publishing,
ANU.

HARLAN, J. 1995. The living field: our agricultural


heritage. Oxford: Oxford University Press.
HARRIS, D.R. 1995. Early agriculture in New Guinea
and the Torres Strait divide, Antiquity 69 (Special
Number 265): 848-54.
1996. The origins and spread of agriculture and
pastoralism in Eurasia: an

2002. Gourds in New Guinea, Asia and the Pacific, in


S. Bedford, C. Sand and D. Burley (eds.), Fifty
years in the field. Essays in honour and celebration of
Richard Shutler Jr.s archaeological career: 69-78.
New Zealand Archaeological Journal Monograph
25, Auckland: Auckland Museum.

(ed.). 1996. The origins and spread of agriculture and


pastoralism in Eurasia. London: UCL Press.
HARRIS, D.R. & G.C. HILLMAN (eds.). 1989. Foraging
and Farming: the evolution of plant exploitation.
London: Unwin Hyman.

GOLSON, J. & P.J. HUGHES. 1980. The appearance of


plant and animal domestication in New Guinea,
Journal de la Socit des Ocanistes 36: 294-303.

HARRIS, E.C. 1977. Hed mound: a New Guinea house


site. Unpublished manuscript on file at RSPAS,
ANU.

GOSDEN, C. & J. HATHER (eds.). 1999. The prehistory of


food: appetites for change. London: Routledge.
GREEN, R. 1991. The Lapita cultural complex: current
evidence and proposed models, Bulletin of the IndoPacific Prehistory Association 11: 295-305.

HARRIS, E.C. & P.J. HUGHES. 1978. An early


agricultural system at Mugumamp Ridge, Western
Highlands Province, Papua New Guinea, Mankind
11 (4): 437-45.

GROUBE, L. 1989. The taming of the rainforests: a


model for Late Pleistocene forest exploitation in
New Guinea, in D.R. Harris and G.C. Hillman
(eds.): 292-304.

HAYWOOD, M. 1985. Soil development in Flandrian


landscapes: River Severn case-study, in J.
Boardman (ed.), Soils and Quaternary landscape
evolution: 281-99. Chichester: John Wiley.

GROUBE, L., J. CHAPPELL, J. MUKE & D. PRICE. 1986. A


40,000 year-old human occupation site at Huon
Peninsula, Papua New Guinea, Nature 324(4):
453-5.

HOPE, G.S. 1983. The vegetational changes of the last


20,000 years at Telefomin, Papua New Guinea,
Singapore Journal of Tropical Geography 4: 25-33.
HOPE, G.S. & J. GOLSON. 1995. Late Quaternary
change in the mountains of New Guinea, Antiquity
69 (Special Number 265): 818-30.

HABERLE, S.G. 1993. Late Quaternary environmental


history of the Tari Basin, Papua New Guinea.
Unpublished PhD thesis, Australian National
University.

855

Research

Tim Denham, Simon Haberle & Carol Lentfer

New evidence and revised interpretations of early agriculture in Highland New Guinea
HOPE, J.H. & G.S. HOPE. 1976. Palaeoenvironments
for man in New Guinea, in R.L. Kirk and A.G.
Thorne (eds.), The Biological Origins of the
Australians: 29-54. Canberra: Australian Institute
of Aboriginal Studies Publication.

PAIN, C.F., C.J. PIGRAM, R.J. BLONG & G.O. ARNOLD.


1987. Cainozoic geology and geomorphology of
the Wahgi Valley, central highlands of Papua New
Guinea, BMR Journal of Australian Geology and
Geophysics 10: 267-75.

HUGHES, P.J., M.E. SULLIVAN & D. YOK. 1991. Human


induced erosion in a Highlands catchment in
Papua New Guinea: the prehistoric and
contemporary records, Zeitschrift fr
Geomorphologie Suppl. 83: 227-39.

POWELL, J.M. 1970a. The history of agriculture in the


New Guinea Highlands, Search 1(5): 199-200.
1970b. The impact of man on the vegetation of the
Mount Hagen region, New Guinea. Unpublished
PhD thesis, Australian National University.

KIRCH, P.V. 1989. Second millenium B.C. aboriculture


in Melanesia: archaeological evidence from the
Mussau Islands, Economic Botany 43(2): 225-40.

1976. Ethnobotany, in K. Paijmans (ed.), New Guinea


vegetation: 106-83. Canberra: CSIRO and ANU
Press.

LEBOT, V. 1999. Biomolecular evidence for plant


domestication in Sahul, Genetic Resources and Crop
Evolution 46: 619-28.

1980. Studies of New Guinea vegetation history,


Proceedings of the IVth International Palynology
Conference, Lucknow (1976-7) 3: 11-20.

LOY, T., M. SPRIGGS & S. WICKLER. 1992. Direct


evidence for human use of plants 28,000 years ago:
starch residues on stone artefacts from northern
Solomon Islands, Antiquity 66: 898-912.

1981. The origins of agriculture in New Guinea,


Proceedings of the IVth International Palynology
Conference, Lucknow (1976-7) 3: 295-310.
1982a. Plant resources and palaeobotanical evidence
for plant use in the Papua New Guinea Highlands,
Archaeology in Oceania 17: 28-37.

MATTHEWS, P.J. 1991. A possible wild type taro:


Colocasia esculenta var. aquatalis, Indo-Pacific
Prehistory Association Bulletin 11: 69-81.

1982b. The history of plant use and mans impact on


the vegetation, in J.L. Gressitt (ed.), Biogeography
and ecology of New Guinea. Volume 1: 207-27. The
Hague: Junk.

1995. Aroids and Austronesians, Tropics 4 (2): 105-26.


2003a. Identification of Benincasa hispida (wax gourd)
from the Kana archaeological site, Western
Highlands Province, Papua New Guinea, in T.P.
Denham and C. Ballard (eds.): 186-91.

1984. Ecological and palaeoecological studies at Kuk


I: below the grey clay. Unpublished manuscript on
file at RSPAS, ANU.

2003b. Taro planthoppers (Tarophagus spp.) in


Australia and the origins of taro (Colocasia
esculenta) in Oceania, in T.P. Denham and C.
Ballard (eds.): 192-202.

POWELL, J.M. & S. HARRISON. 1982. Haiyapugua:


aspects of Huli subsistence and swamp cultivation..
Port Moresby: Department of Geography, UPNG
(Occasional Paper No. 1, New Series).

MBIDA, C.M., H. DOUTRELEPONT, L. VRYDAGHS, R.L.


SWENNEN, R.J. SWENNEN, H. BEECKMAN, E. DE
LANGHE & P. DE MARET. 2001. First archaeological
evidence of banana cultivation in central Africa
during the third millennium before present,
Vegetation History and Archaeobotany 10: 1-6.

POWELL, J.M., A. KULUNGA, R. MOGE, C. PONO, F.


ZIMIKE & J. GOLSON. 1975. Agricultural traditions
in the Mount Hagen area. Port Moresby:
Department of Geography, UPNG (Occasional
Paper No. 12).

MBIDA, C.M., W. VAN NEER, H. DOUTRELEPONT & L.


VRYDAGHS. 2000. Evidence for banana cultivation
and animal husbandry during the first millennium
BC in the forest of Southern Cameroon, Journal of
Archaeological Science 27: 151-62.

RICHERSON, P.J., R. BOYD & R.L. BETTINGER. 2001.


Was agriculture impossible during the Pleistocene
but mandatory during the Holocene? A climate
change hypothesis, American Antiquity 66(3): 387411.

MCALPINE, J.R., G. KEIG & R. FALLS. 1983. Climate of


Papua New Guinea. Canberra: CSIRO and ANU
Press.

ROSCOE, P. 2002. The hunters and gatherers of New


Guinea, Current Anthropology 43(1): 153-62.
SAUER, C.O. 1952. Agricultural origins and their
dispersals. New York: American Geographical
Society.

MOUNTAIN, M.J. 1991. Highland New Guinea huntergatherers: the evidence of Nombe Rockshelter, Simbu,
with emphasis on the Pleistocene. Unpublished PhD
thesis, Australian National University.

SIMMONDS, N.W. 1962. The evolution of the bananas.


London: Longmans.

MUKE, J. & H. MANDUI. 2003. Shadows of Kuk:


evidence for prehistoric agriculture at Kana, Wahgi
Valley, Papua New Guinea, in T.P. Denham and C.
Ballard (eds.): 177-85.

SMITH, B.D. 1998. The emergence of agriculture. New


York: Scientific American Library.

856

SPRIGGS, M. 1996. Early agriculture and what went


before in Island Melanesia: continuity or intrusion?
in D.R. Harris (ed.): 524-37.

TERRELL, J.E. & R. WELSCH. 1997. Lapita and the


temporal geography of prehistory, Antiquity 71:
548-72.

2001. Who cares what time it is? The importance of


chronology in Pacific archaeology, in A. Anderson,
I. Lilley and S. OConnor (eds.), Histories of old
ages: essays in honour of Rhys Jones: 237-49.
Canberra: Pandanus Books, Australian National
University.

WILSON, S.M. 1985. Phytolith evidence from Kuk, an


early agricultural site in New Guinea, Archaeology
in Oceania 20: 90-7.

STEVENS, M.L., R.M. BOURKE & B.R. EVANS (eds.).


South Pacific indigenous nuts. Canberra: ACIAR.

YEN, D.E. 1991. Polynesian cultigens and cultivars: the


questions of origin, in P.A. Cox and S.A. Banack
(eds.), Islands, plants and Polynesians: an
introduction to Pacific ethnobotany: 67-95.
Portland: Dioscorides Press.

STUIVER, M & P.J. REIMER. 1993. Extended 14C


database and revised CALIB radiocarbon
calibration program, Radiocarbon 35:215-230.

1995. The development of Sahul agriculture with


Australia as bystander, Antiquity 69, Special
Number 265: 831-47.

SWADLING, P., N. ARAHO & B. IVUYO. 1991. Settlements


associated with the inland Sepik-Ramu Sea,
Bulletin of the Indo-Pacific Prehistory Association 11:
92-110.

1996. Melanesian aboriculture: historical perspectives


with emphasis on the genus Canarium, in M.L.
Stevens, R.M. Bourke and B.R. Evans (eds.): 3644.

TERRELL, J.E. 2002. Tropical agroforestry, coastal


lagoons and Holocene prehistory in Greater Near
Oceania, in Y. Shuji and P.J. Matthews (eds.),
Proceedings of the International Area Studies
Conference VII: vegeculture in Eastern Asia and
Oceania: 195-216. Osaka: National Museum of
Ethnology.

1998. Subsistence to commerce in Pacific agriculture:


some four thousand years of plant exchange, in
H.D.V. Pendergast, N.L. Etkin, D.R. Harris and
P.J. Houghton (eds.), Plants for food and medicine:
161-83. Kew: Royal Botanic Gardens.

TERRELL, J.E., J.P. HART, S. BARUT, N. CELLINESE, A.


CURET, T.P. DENHAM, H. HAINES, C.M. KUSIMBA,
K. LATINIS, R. OKA, J. PALKA, M.E.D. POHL, K.O.
POPE, J.E. STALLER & P.R. WILLIAMS. 2003.
Domesticated landscapes: the subsistence ecology
of plant and animal domestication, Journal of
Archaeological Method and Theory 10(4): 323-68.

857

Research

Tim Denham, Simon Haberle & Carol Lentfer

You might also like