07.02.2013 Views

chapter 1 - ResearchSpace@Auckland - The University of Auckland

chapter 1 - ResearchSpace@Auckland - The University of Auckland

chapter 1 - ResearchSpace@Auckland - The University of Auckland

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Copyright Statement<br />

<strong>The</strong> digital copy <strong>of</strong> this thesis is protected by the Copyright Act 1994 (New<br />

Zealand). This thesis may be consulted by you, provided you comply with the<br />

provisions <strong>of</strong> the Act and the following conditions <strong>of</strong> use:<br />

• Any use you make <strong>of</strong> these documents or images must be for<br />

research or private study purposes only, and you may not make<br />

them available to any other person.<br />

• Authors control the copyright <strong>of</strong> their thesis. You will recognise the<br />

author's right to be identified as the author <strong>of</strong> this thesis, and due<br />

acknowledgement will be made to the author where appropriate.<br />

• You will obtain the author's permission before publishing any<br />

material from their thesis.<br />

To request permissions please use the Feedback form on our webpage.<br />

http://researchspace.auckland.ac.nz/feedback<br />

General copyright and disclaimer<br />

In addition to the above conditions, authors give their consent for the digital<br />

copy <strong>of</strong> their work to be used subject to the conditions specified on the Library<br />

<strong>The</strong>sis Consent Form


READING LAPITA IN NEAR OCEANIA:<br />

INTERTIDAL AND SHALLOW-WATER POTTERY<br />

SCATTERS, ROVIANA LAGOON, NEW GEORGIA,<br />

SOLOMON ISLANDS<br />

MATTHEW WALTER FELGATE<br />

A thesis submitted in partial fulfilment <strong>of</strong> the requirements for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy in Anthropology, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>, 2003


ABSTRACT<br />

Lapita is the name given by archaeologists to a material culture complex distributed from<br />

Papua New Guinea to Samoa about 3000 years ago, which marks major economic changes<br />

in Near Oceania and the first settlement by humans <strong>of</strong> Remote Oceania. Those parts <strong>of</strong><br />

Solomon Islands that lie in Near Oceania, together with Bougainville, comprise a large gap<br />

in the recorded distribution <strong>of</strong> Lapita, which the current research seeks to explain. At<br />

Roviana Lagoon, centrally located in this gap, scatters <strong>of</strong> pottery, stone artefacts, and other<br />

stone items are found in shallow water in this sheltered, landlocked lagoon, initially thought<br />

to be late derivatives <strong>of</strong> Lapita. This research seeks method and theory to aid in the<br />

interpretation <strong>of</strong> this type <strong>of</strong> archaeological record.<br />

Intensive littoral survey discovered a wider chronological range <strong>of</strong> pottery styles<br />

than had previously been recorded, including materials attributable directly to the Lapita<br />

material culture complex. A study <strong>of</strong> vessel brokenness and completeness enabled sample<br />

evaluation, estimation <strong>of</strong> a parent population from which the sample derived, assessment<br />

<strong>of</strong> the state <strong>of</strong> preservation <strong>of</strong> the sample, and systematic choice <strong>of</strong> unit <strong>of</strong> quantification.<br />

Studies <strong>of</strong> wave exposure <strong>of</strong> collection sites and taphonomic evidence from sherds<br />

concluded that the cultural formation process <strong>of</strong> these sites was stilt house settlement (as<br />

found elsewhere in Near Oceania for Lapita) over deeper water than today. Falling relative<br />

sea levels and consequent increasing effects <strong>of</strong> swash-zone processes have resulted in high<br />

archaeological visibility and poor state <strong>of</strong> preservation at Roviana Lagoon.<br />

Analysis <strong>of</strong> ceramic and lithic variability and spatial analysis allowed the<br />

construction <strong>of</strong> a provisional chronology in need <strong>of</strong> further testing. Indications are that<br />

there is good potential to construct a robust, high-resolution ceramic chronology by<br />

focussing on carefully controlled surface collection from this sort <strong>of</strong> location, ceramic<br />

seriation and testing/calibration using direct dating by AMS radiocarbon and<br />

<strong>The</strong>rmoluminescence.<br />

Data on preservation and archaeological visibility <strong>of</strong> stilt house settlements along<br />

a sheltered emerging coastline allows preservation and visibility for this type <strong>of</strong> settlement<br />

to be modeled elsewhere. When such a model is applied to other areas <strong>of</strong> the Lapita gap,<br />

which are predominantly either less favourable for preservation or less favourable for<br />

archaeological visibility, the gap in the distribution <strong>of</strong> Lapita can be seen to be an area <strong>of</strong><br />

low probability <strong>of</strong> detection by archaeologists, meaning there is currently no evidence for<br />

absence <strong>of</strong> settlement in the past, and good reason to think that Lapita was continuously<br />

distributed across Near Oceania as a network <strong>of</strong> stilt village settlement. This finding<br />

highlights the need for explicit models <strong>of</strong> probability <strong>of</strong> detection to discover or read the<br />

Lapita archaeological record.<br />

Keywords: pottery; Lapita; formation processes; surface archaeology; tidal archaeology;<br />

Oceania


ACKNOWLEDGEMENTS<br />

Primary supervisor, Dr. Peter Sheppard, got the New Georgia Archaeological Survey <strong>of</strong>f<br />

the ground (and into the sea), an achievement that made this research possible. As a<br />

supervisor he gave me a lot <strong>of</strong> rope, as well as the occasional well-timed tug back to the<br />

topic, and has always been available when needed. Peter was instrumental in obtaining<br />

institutional financial support in the form <strong>of</strong> a <strong>University</strong> <strong>of</strong> <strong>Auckland</strong> Doctoral<br />

Scholarship, and a research stipend from the Marsden Fund. Field expenses and substantial<br />

analytical expenses were funded primarily from grants to the NGAS by the <strong>University</strong> <strong>of</strong><br />

<strong>Auckland</strong>, National Geographic Research and the Marsden Fund. Peter has also been a<br />

steady companion and mentor in the field and at all stages <strong>of</strong> the research, and he and<br />

Debbie have made my family and I welcome in their home on numerous occasions.<br />

This research in effect continues an academic program begun in the 1970s by<br />

Pr<strong>of</strong>essor Roger C. Green, as the Southeast Solomons Culture History Project, which<br />

eventually brought Peter Sheppard to this part <strong>of</strong> the world to work on Lapita lithics from<br />

Solomon Islands. It was Roger Green who first focused my attention on Oceanic<br />

prehistory, initially with an ample shirt emblazoned with pacific canoes, and later through<br />

an intense series <strong>of</strong> lectures on Oceanic Prehistory, capped by his last 3-week fieldschool<br />

before retirement from active fieldwork in the mid 1980s. Roger predicted in 1978 that<br />

Lapita would be found at New Georgia, and like Peter Sheppard, has been a mentor<br />

throughout. He also organized and personally funded petrographic analysis <strong>of</strong> ceramic<br />

tempers, and made his extensive personal library available.<br />

Dr. Simon Holdaway generously accepted the task <strong>of</strong> secondary supervision in midresearch<br />

when the rules changed, and has, through suggestions in the lab, comments on<br />

drafts, and by convening an archaeology theory reading group, contributed substantially<br />

to the outcome.<br />

Material and moral support from family were essential to successful completion <strong>of</strong><br />

the research, especially from my parents, who largely financed the final years, and elder<br />

brother, who was always ready to help pack up the house and store incredible<br />

accumulations <strong>of</strong> household effects every time we decamped to the Solomons, and is still<br />

doing so.<br />

In Solomon Islands, support and encouragement from Mr Lawrence Foanoata,<br />

Director <strong>of</strong> the National Museum has been constant and essential. Permissions from the<br />

Research committee <strong>of</strong> the Ministry <strong>of</strong> Education, and from the Western Province<br />

Administration were key to the success <strong>of</strong> the research. Mr John Keopo <strong>of</strong> the National<br />

Museum was instrumental in the smooth running <strong>of</strong> the first field season. Rhys and<br />

Margaret Richards provided hospitality in Honiara on more than one occasion, and Rhys<br />

was instrumental as the New Zealand High Comissioner in obtaining a residence permit<br />

during extended fieldwork.<br />

II


Mr Kenneth Roga, Western Province Field Archaeologist, has been a primary architect <strong>of</strong><br />

fieldwork throughout. Unflagging energy in the field, his encyclopaedic network <strong>of</strong> wantok<br />

which has ensured a warm welcome on so many occasions, commitment to archaeology<br />

despite difficult political circumstances, and a range <strong>of</strong> talents too numerous to mention<br />

have been invaluable. A rock-steady figure at the helm <strong>of</strong> a canoe over many long journeys<br />

through sometimes dangerous seas, his and Janet Roga’s hospitality in Gizo on so many<br />

occasions is deeply appreciated.<br />

At Roviana Lagoon, permission for research was received from the late Chief<br />

Johnathan Roni. Chief Joseph Kama <strong>of</strong> Kaliquongu and Chief Nathan Kera <strong>of</strong> Saikile each<br />

consented to research in areas under their jurisdiction. Thanks are also due to Mr. Solomon<br />

Roni and his family, especially all at Miho in Sasavele: Sam Roni and Waring, Martha<br />

Roni, Aggie Roni and Abel Kae, my father-in-law Sale Maebule and late mother-in-law<br />

Mabelo Roni, Sitiveni, my wife Noseduri, and many <strong>of</strong> the next generation too numerous<br />

to mention. Thanks also are due to members <strong>of</strong> the Elelo community who helped out in so<br />

many ways during our extended stay, especially Hetti Lanni, Barrie and Selina Ford and<br />

George and Visa Sapolo. <strong>The</strong> late Mr Phillip Lanni generously assisted with his time during<br />

a visit to Gharanga, despite ill health. In Munda, thanks are due to David Kera and family,<br />

Dave and Mariana Cook, and Trevor and Zahi Cumberland, for assistance with<br />

communications and travel on many occasions. In Saikile, John Kororo and family abetted<br />

us on our search for pottery at Mbaraulu, and Sae Oka was a prime mover in the survey<br />

for pottery on Ndora Island.<br />

During analysis and writeup at <strong>Auckland</strong> many individuals contributed their skills:<br />

Dilys Johns and Dr Rod Wallace supplied technical and conservation advice, Hamish Mc<br />

Donald and Tim Mackrell took all the artefact photographs (or at least all the good ones),<br />

Joan Lawrence illustrated the pottery and lithic artefacts, Dr Simon Bickler collaborated<br />

on the problem <strong>of</strong> sample interpretation, and wrote a simulation program to estimate<br />

parent population <strong>of</strong> vessels. Dr Robin Parker <strong>of</strong> the Geology Department provided thin<br />

section descriptions <strong>of</strong> lithics. Barry Curnham, also in Geology, showed me how to make<br />

epoxy-impregnated ceramic thin sections. Peter Sheppard examined chert thin sections. Bill<br />

Dickinson <strong>of</strong> the <strong>University</strong> <strong>of</strong> Arizona analysed and reported on ceramic thin sections, and<br />

collaborated on a paper writing up the results. Jim Feathers <strong>of</strong> the Anthropology<br />

Department, <strong>University</strong> <strong>of</strong> Washington, did the thermoluminescence analysis and was<br />

generous with his time in discussing progress and results. Dr Christine Prior <strong>of</strong> the Rafter<br />

Radiocarbon Lab organized AMS dating <strong>of</strong> samples from Honiavasa and Hoghoi, including<br />

some careful experimentation with the Honiavasa sample. Pr<strong>of</strong>essor Jim Allen gave<br />

permission to cite the 1984 and 1985 Lapita Homeland Project Field Reports.<br />

Discussions with Dr Stuart Bedford, Moira Doherty, Dr Simon Best and Dr Simon Bickler,<br />

contributed to the research and identified several errors. Stuart, Simon Best and Moira<br />

provided useful comments on parts <strong>of</strong> the manuscript. Stuart and Carolyn made me<br />

welcome in their home during the final stages <strong>of</strong> writing.<br />

III


CHAPTER 1:<br />

CHAPTER 2:<br />

Contents:<br />

RESEARCH QUESTIONS AND METHODOLOGY ................ 1<br />

Introduction: ................................................. 1<br />

<strong>The</strong> Research Region: .......................................... 6<br />

<strong>The</strong> Roviana Early Ceramic Archaeological Record: ............... 14<br />

Research Questions: .......................................... 15<br />

Approaches to Ceramic Classification and Analysis: ................ 21<br />

Quantification: ............................................... 34<br />

Seriation <strong>The</strong>ory : a Review: ................................... 43<br />

Form and Function: a Review : ................................. 57<br />

Chapter summary and conclusions: .............................. 70<br />

A REVIEW OF CONSTRUCTIONS OF LAPITA TEMPORAL<br />

VARIABILITY .............................................. 77<br />

Introduction: ................................................ 77<br />

Green’s 1978 Lapita Ceramic Series: ............................ 79<br />

Anson’s Early Far Western Lapita: .............................. 85<br />

Mussau: ..................................................... 88<br />

<strong>The</strong> “Changing” Face <strong>of</strong> Lapita: ................................ 89<br />

Summerhayes, West New Britain, Anir, and a Three-stage Lapita Ceramic<br />

Series: ................................................ 93<br />

<strong>The</strong> Watom Lapita Series: .....................................103<br />

Wahome’s Seriation <strong>of</strong> Admiralties Pottery Assemblages: ...........107<br />

Specht on Buka: ..............................................109<br />

Wickler on Buka: .............................................111<br />

Vanuatu: ....................................................116<br />

New Caledonia: ..............................................119<br />

<strong>The</strong> View from the East: Fiji and Tonga: .........................123


Lapita Temporal Variability -Conclusions: ...................... 125<br />

CHAPTER 3:<br />

CHAPTER 4:<br />

CHAPTER 5:<br />

SCALE AND METHOD OF FIELD SURVEY REQUIRED TO GENERATE<br />

A SAMPLE OF LAPITA SITES FOR SERIATION IN THE NEW<br />

GEORGIA REGION ......................................... 133<br />

Introduction: ............................................... 133<br />

Review <strong>of</strong> Near-oceanic Survey Methods and Results: ............. 139<br />

Sampling <strong>The</strong>ory and the Roviana/Kaliquongu Survey Regions: ..... 158<br />

Survey Methods: ............................................ 163<br />

Results: .................................................... 188<br />

Discussion and Conclusions: the Two Surveys. .................... 190<br />

CERAMICS: UNITS OF DESCRIPTION ....................... 193<br />

Introduction: ............................................... 193<br />

Database Structure: .......................................... 195<br />

Thickness, Form and Decoration <strong>of</strong> Various Vessel Parts Represented on<br />

Sherds: .............................................. 202<br />

Vessel Form: ............................................... 205<br />

Decoration: ................................................. 219<br />

Transforming Relational Data into a Flat Table: .................. 242<br />

Chapter Summary and Conclusions: ............................ 244<br />

SAMPLE EVALUATION AND QUANTIFYING SAMPLE SIZE. ... 247<br />

Introduction ................................................ 247<br />

Methods for Establishing Vessel Brokenness and Completeness: ..... 248<br />

Simulation Approach: ........................................ 253<br />

Statistical Approaches: ....................................... 257<br />

Simulation Results: .......................................... 259<br />

Statistical Results: ........................................... 262


Discussion: ..................................................263<br />

CHAPTER 6:<br />

CHAPTER 7:<br />

CHAPTER 8:<br />

Summary and Conclusions: ....................................263<br />

WAVE EXPOSURE ..........................................269<br />

Introduction: ................................................269<br />

Review: .....................................................270<br />

Roviana Intertidal Ceramics and Waves: .........................272<br />

Solomon Islands Tides: ........................................277<br />

Winds: ......................................................280<br />

Measuring Fetch: .............................................282<br />

Results: .....................................................283<br />

Discussion: ..................................................283<br />

Chapter Summary and Conclusions: ......................286<br />

SITE/ASSEMBLAGE FORMATION PROCESSES (SHERD<br />

TAPHONOMY). .............................................289<br />

Introduction: ................................................289<br />

Review: Middle Range <strong>The</strong>ory for the Identification <strong>of</strong> Formation Processes:<br />

......................................................291<br />

Models <strong>of</strong> Cultural Formation Process: ...........................298<br />

Models <strong>of</strong> Post-depositional Formation Process: ...................299<br />

Summary <strong>of</strong> Models: ..........................................301<br />

Qualitative Evidence From Samples: .............................302<br />

Quantitative Evidence: Size/Density Effects and the Structure and Context<br />

<strong>of</strong> the Deposit: ..........................................305<br />

Sea Level History: ............................................312<br />

Chapter Summary and Conclusions: .............................313


VESSEL FORM VARIABILITY AND VESSEL FUNCTION ....... 319<br />

CHAPTER 9:<br />

Introduction: ............................................... 319<br />

Form: Initial Nominal Classification: ........................... 322<br />

Metric Variability Within Form 6: ............................. 331<br />

Inferring Vessel Form from Body Sherds: ....................... 339<br />

Vessel Function from Evidence for Use-Alteration: ............... 346<br />

Secondary Smoothing <strong>of</strong> Vessel Interiors, and Interior Neck Form: . . 347<br />

Shoulder Form Classes <strong>of</strong> Form 6 Vessels: ...................... 351<br />

Sherd Restriction Factor: ..................................... 353<br />

Chapter Summary: .......................................... 355<br />

CERAMIC DECORATIVE CLASSIFICATION AND<br />

DECORATION/FORM VARIABILITY: ....................... 359<br />

Introduction: ............................................... 359<br />

Classification <strong>of</strong> Linear Motifs: ................................ 360<br />

Part Representation in the Potsherd Sample: ..................... 361<br />

Covariation Between Vessel Part and Decoration: ................. 364<br />

Summary <strong>of</strong> Decorative Structure Across the Vessel: .............. 371<br />

Interpretation <strong>of</strong> Decorative Co-variation by Vessel Part: .......... 373<br />

Attribute Frequencies: ....................................... 374<br />

Variability <strong>of</strong> Impressed Lips: ................................. 375<br />

Decorative Attributes and Vessel Form: ......................... 383<br />

Chapter Summary: .......................................... 391<br />

CHAPTER 10:<br />

LITHICS ................................................... 395<br />

Introduction: ............................................... 395<br />

Review: .................................................... 398<br />

Roviana Lithic Artefacts: ..................................... 401<br />

Chert Flakes/Fragments: ..................................... 411<br />

Analysis <strong>of</strong> Hoghoi Water-rounded and Fractured Volcanic Manuports:


......................................................412<br />

Chapter Summary: ...........................................417<br />

CHAPTER 11:<br />

INTRASITE SPATIAL STRUCTURE OF CERAMIC AND LITHIC<br />

VARIABILITY: ..............................................421<br />

Introduction: ................................................421<br />

Selection <strong>of</strong> Sites for Intrasite Spatial Analysis: ....................423<br />

Objectives: ..................................................423<br />

Method .....................................................425<br />

Zangana: ....................................................428<br />

Hoghoi: .....................................................438<br />

Honiavasa: ..................................................443<br />

Chapter Summary and Conclusions. .............................447<br />

CHAPTER 12:<br />

CHRONOLOGY .............................................451<br />

Introduction: ................................................451<br />

AMS Radiocarbon Dates on Potsherds: ..........................453<br />

<strong>The</strong>rmoluminescence (TL) Dates from Quartz-Calcite Sherds: .......458<br />

Roviana TL Data: ............................................463<br />

Seriation: ....................................................465<br />

Discussion <strong>of</strong> Correspondence Analyses: the Alternatives: ...........478<br />

Conclusions: Integrating 14 C, TL and Seriation: ...................480<br />

CHAPTER 13:<br />

SUMMARY AND CONCLUSIONS .............................483<br />

Introduction: ................................................483<br />

Summary: ...................................................484<br />

External Comparisons: ........................................498<br />

<strong>The</strong> Lapita Gap as an Area <strong>of</strong> Low Probability <strong>of</strong> Detection : ........503<br />

Intertidal-Zone and Shallow-Water Archaeology: ..................504


REFERENCES .................................................... 509


List <strong>of</strong> Tables:<br />

Table 1: Sampling, assemblage richness, and the Jaccard coefficient (p=present, a=absent).<br />

............................................................ 52<br />

Table 2: Reef/ Santa Cruz motif counts as given in Anson 1983. .............. 80<br />

Table 3: Sherd counts and MNI assembled from various tables in Parker (1981). . 81<br />

Table 4: Relative proportions <strong>of</strong> dentate and incised, recalculated from data presented by<br />

Donovan (1973). .............................................. 82<br />

Table 5: Summary <strong>of</strong> a review <strong>of</strong> survey methods and Lapita results. ...........155<br />

Table 6: Site densities by period for Roviana and Kaliquongu surveys. .........188<br />

Table 7: Structure <strong>of</strong> master record for each sherd. .........................196<br />

Table 8: Classes <strong>of</strong> mineral identified at 10x magnification in reflected light. ....199<br />

Table 9: Examples <strong>of</strong> descriptive syntax for tempers. .......................200<br />

Table 10: Temper groupings after Dickinson 2000. .........................201<br />

Table 11: Data structure <strong>of</strong> table <strong>of</strong> thicknesses for each part <strong>of</strong> the sherd. .......204<br />

Table 12: Data structure for the table <strong>of</strong> records <strong>of</strong> sherd form attributes by vessel part.<br />

............................................................206<br />

Table 13: Data structure for decoration records.............................220<br />

Table 14: Decorative techniques. .......................................221<br />

Table 15: Decoration pattern definitions..................................229<br />

Table 16: Decorative elements..........................................240<br />

Table 17: Data structure for flat table <strong>of</strong> summary data <strong>of</strong> sherd properties. ......245<br />

Table 18: Variation in lip brokenness between sites. ........................250<br />

Table 19: Selection <strong>of</strong> sample for estimating vessel completeness. .............252<br />

Table 20: Breakage population estimate for the combined Roviana highly decorated lip<br />

sample using the statistic <strong>of</strong> Chao 1984. ............................262<br />

Table 21: Fetch measurements for collection sites as an indicator <strong>of</strong> wave exposure.<br />

............................................................283<br />

Table 22: Total sherd count and average sherd area by collection site, resulting from<br />

combined effects <strong>of</strong> collection intensity and wave exposure. ............286<br />

Table 23: Ratio <strong>of</strong> lip-rim sherds to body sherds as a measure <strong>of</strong> collector effect.


........................................................... 309<br />

Table 24: Ratios <strong>of</strong> plain to decorated sherds, controlled by vessel part (lips, rims, necks<br />

and shoulders. ............................................... 310<br />

Table 25: Body sherd to neck sherd ratio as an index <strong>of</strong> sherd skipping. ........ 311<br />

Table 26: Counts <strong>of</strong> vessel form classes. ................................ 336<br />

Table 27: <strong>The</strong> attribute “even/not even” by site. .......................... 348<br />

Table 28: Relative abundance <strong>of</strong> “hard neck” interior pr<strong>of</strong>iles to “not hard”. .... 350<br />

Table 29: Relative abundance <strong>of</strong> even interiors with hard neck interior pr<strong>of</strong>iles to even<br />

interiors without hard pr<strong>of</strong>ile. ................................... 350<br />

Table 30: Sherd restriction factor comparison <strong>of</strong> site samples. ............... 353<br />

Table 31: Preservational bias <strong>of</strong> vessel parts as an explanation for differences in sherd<br />

restriction factor (data has the form average thickness (count) standard deviation).<br />

........................................................... 354<br />

Table 32: Part representation, all sherds including necked sherds but excluding carinated<br />

and/or inverted-rim sherds. ..................................... 362<br />

Table 33: Part representation for inverted-rim sherds (no neck corner point expected from<br />

morphology <strong>of</strong> sherd). ......................................... 362<br />

Table 34: Carinated sherds (excluding carinated sherds with inverted rims). .... 363<br />

Table 35: Form strength variation for different pottery styles and the effect on part<br />

representation. ............................................... 363<br />

Table 36: Lip decoration and rim decoration.............................. 365<br />

Table 37: Lip decoration and neck decoration............................. 366<br />

Table 38: Cross-tabulation <strong>of</strong> lip decoration with shoulder decoration. ......... 367<br />

Table 39: Cross-tabulation <strong>of</strong> rim decoration and neck decoration. ............ 368<br />

Table 40: Cross-tabulation <strong>of</strong> rim decoration and shoulder decoration. ......... 369<br />

Table 41: Cross-tabulation <strong>of</strong> neck decoration and shoulder decoration. ........ 370<br />

Table 42: Counts <strong>of</strong> occurrences <strong>of</strong> decorative attributes. ................... 375<br />

Table 43: Petrographic descriptions <strong>of</strong> lithic artefacts. ...................... 407<br />

Table 44: Petrographic classification <strong>of</strong> manuports collected at Hoghoi. ....... 414<br />

Table 45: Size comparisons between petrographic classes <strong>of</strong> lithic manuports. . . 415<br />

Table 46: Radiocarbon determinations from pottery (calibrated using OxCal 3.5, Stuiver<br />

et al. 1998 atmospheric data). ................................... 454


Table 47: <strong>The</strong>rmoluminescence dating results. Precision shown is at confidence limits <strong>of</strong><br />

one standard deviation. .........................................464<br />

Table 48: Definitions <strong>of</strong> attribute codes used in seriation tables and plots. .......465<br />

Table 49: Relative contribution <strong>of</strong> first and second components <strong>of</strong> CA using all attributes<br />

and forms. ...................................................468<br />

Table 50: CA diagnostics by attribute, all attributes included. .................468<br />

Table 51: CA diagnostics by site, all attributes included (*=inertia outlier). ......470<br />

Table 52: Eigenvalues and contributions to intertia <strong>of</strong> CA components, for data excluding<br />

form-correlated attributes. .......................................473<br />

Table 53: CA diagnostic table for attributes, form-correlated attributes omitted (*=intertia<br />

outlier). ......................................................473<br />

Table 54: CA diagnostics by site, form-correlated attributes omitted (*=inertia outlier).<br />

............................................................475<br />

Table 55: CA diagnostics for attributes, omitting Honiavasa data and form-correlated<br />

attributes. ....................................................477<br />

Table 56: CA diagnostics by site, omitting Honiavasa data and form-correlated attributes.<br />

............................................................477<br />

Table 57: CA eigenvalues and component contributions to inertia omitting Honiavasa data<br />

and form-correlated attributes ....................................477<br />

Table 58: Summary <strong>of</strong> variability .......................................479


List <strong>of</strong> Figures:<br />

Figure 1: Near Oceania, Remote Oceania and Solomon Islands, showing the location <strong>of</strong><br />

the New Georgia Group. ......................................... 3<br />

Figure 2: Map <strong>of</strong> New Georgia Group showing principal geological formations (after<br />

Coulson, Dunkerly, Hughes and Ridgeway 1987). ..................... 9<br />

Figure 3: <strong>The</strong> author providing scale for solution notches in Plio-Pleistocene limestone<br />

cliff near Saikile passage, Roviana Lagoon (photograph courtesy <strong>of</strong> Peter<br />

Sheppard) .................................................... 14<br />

Figure 4: Paniavile at low tide, looking north, with several inhabited islets and the New<br />

Georgia mainland in the distance. ................................. 16<br />

Figure 5: Processes <strong>of</strong> formation <strong>of</strong> the Archaeological record (After Felgate and Bickler<br />

n.d., adapted from De Boer 1983): inferring a breakage population from an<br />

archaeological sample........................................... 44<br />

Figure 6: Roviana and Kaliquongu surveys as samples <strong>of</strong> the New Georgia lagoon and<br />

barrier island system. ...........................................158<br />

Figure 7: Slab-built carinated vessels from Honiavasa initially assigned to the Lapita<br />

period (this assignment is examined in more detail in Chapters 8 and 9): the solid<br />

vertical line in HV.2.464 represents an estimated location <strong>of</strong> the central vertical<br />

avis (CVA) <strong>of</strong> the pot. ..........................................167<br />

Figure 8: Carinated vessels from Honiavasa, showing double-line markes on some design<br />

zones, bands <strong>of</strong> nubbins at the neck in two cases, and a band <strong>of</strong> fingernail<br />

impression in one case (top). .....................................168<br />

Figure 9: HV.4.202 has an incised motif laid out in double lines; HV.1.314 is dentate-<br />

stamped, with similar design structure and a double carination; HV.2.341 is<br />

carinated, with the design laid out in applied fillets, bounded horizontally by<br />

decorated lap joins between slabs; HV.2.297 and HV.4.379 have bands <strong>of</strong> single<br />

fingernail impressions and nubbins at the neck respectively. ............169<br />

Figure 10: Incised rims with opposed-pinch fingernail impressed band at the neck,<br />

diagnostic <strong>of</strong> the Miho subgroup <strong>of</strong> Post-Lapita styles. ................171<br />

Figure 11: Miho-style post-Lapita sherds. ................................172<br />

Figure 12: Miho-style post-Lapita sherds with incised shoulders. ..............172


Figure 13: Miho-style post-Lapita sherds with CVA measurements shown (MH290 was<br />

measured at to locations on the pr<strong>of</strong>ile; at the interior <strong>of</strong> the neck orifice and the<br />

exterior shoulder). ............................................ 173<br />

Figure 14: Miho-style post-Lapita sherds; dashed CVA lines are measurements based on<br />

non-circular (uneven) curvature at the pr<strong>of</strong>ile locations indicated by arrows.<br />

........................................................... 174<br />

Figure 15: Gharanga-style post-Lapita sherds. (Gharanga is a short-rim subgroup <strong>of</strong><br />

Gharanga-Kopo which may have multiple bands <strong>of</strong> opposed-pinch fingernail<br />

impression on the shoulder. ..................................... 175<br />

Figure 16: Gharanga-style post-Lapita sherds. ............................ 175<br />

Figure 17: Gharanga/Kopo style post-Lapita sherds (top) and a large Gharanga-style<br />

sherd (bottom) (four measurement points used as arrowed to estimate CVA.<br />

........................................................... 176<br />

Figure 18: Intermediate between Gharanga and Kopo styles: all post-Lapita. .... 177<br />

Figure 19: Kopo-style post-Lapita rim sherds. ............................ 177<br />

Figure 20: Another large Gharanga-style post-Lapita sherd from an unrestricted vessel<br />

form........................................................ 178<br />

Figure 21: Gharanga-style small-orifice post-Lapita sherd showing rolled rim and thin wall<br />

common to this style. .......................................... 178<br />

Figure 22: Kopo-Style post-Lapita sherd (taller, less everted rim than Gharanga style)<br />

from a large-orifice vessel, with bands <strong>of</strong> impressions along both inner and outer<br />

edges <strong>of</strong> the lip. .............................................. 178<br />

Figure 23: Less decorated variant <strong>of</strong> Gharanga-Kopo post-Lapita style, without a strong<br />

corner point in vertical section at the neck. ......................... 179<br />

Figure 24: Large-orifice Gharanga/Kopo-style sherd with deformation <strong>of</strong> the lip into a<br />

wave pattern. ................................................ 179<br />

Figure 25: Large Gharanga-style post-Lapita sherd with typical decoration, including a<br />

band <strong>of</strong> impressions along the inner edge <strong>of</strong> the lip. .................. 179<br />

Figure 26: Gharanga/Kopo-style post-Lapita vessels: most are weakly restricted at the<br />

neck, with short, heavily everted rims. Punctate band at the neck is the most<br />

common decoration in this group, while multiple bands <strong>of</strong> fingernail pinch are<br />

common on the short-rim examples. One sherd (MH.33) had exotic quartz-calcite


hybrid temper (see Chapter 4). ...................................180<br />

Figure 27: Locations mentioned in the text in relation to Roviana and Kaliquongu surveys.<br />

............................................................183<br />

Figure 28: Kaliquongu survey transects: the unfilled symbols represent sites discovered<br />

by informant-prospection during the Roviana survey. .................189<br />

Figure 29: Data structure for each sherd record; each sherd can have many records in the<br />

detail tables pertaining to the various parts <strong>of</strong> the vessel represented. .....195<br />

Figure 30: Major vessel form variants showing part terminology; L=lip, R=rim, N=neck,<br />

S=shoulder, C=carination, B=body; inverted or unrestricted vessels have no neck,<br />

while for restricted vessels with everted rims (the first seven) the only distinction<br />

in neck types is between the double neck (top centre) and the single neck<br />

(including all unlabelled). .......................................203<br />

Figure 31: Lip form variants showing database codes .......................207<br />

Figure 33: Measurement <strong>of</strong> rim depth, rim Vcurve, neck angle and neck Vcurve at<br />

different levels <strong>of</strong> brokenness. ....................................212<br />

Figure 34: Shoulder form measurement. ..................................213<br />

Figure 35: Derivation <strong>of</strong> conical/cannister ( C ) and spheroidal (G) sherds from various<br />

body forms. ..................................................216<br />

Figure 36: Form codes for vessel interiors. ...............................217<br />

Figure 37: Possible conical base sherds from robust vessels. ..................218<br />

Figure 38: Examples <strong>of</strong> applied decoration. ...............................223<br />

Figure 39: Examples <strong>of</strong> applied decoration. ...............................223<br />

Figure 40: Applied decoration on compound rims. .........................224<br />

Figure 41: Examples <strong>of</strong> deformation <strong>of</strong> the lip into a discontinuous band. .......224<br />

Figure 42: Horizontal deformation <strong>of</strong> the lip into a continuous band in a wave pattern.<br />

............................................................225<br />

Figure 43: Examples <strong>of</strong> discontinuous deformation. ........................225<br />

Figure 44: Excision <strong>of</strong> outer lip to form a band <strong>of</strong> notches. ...................226<br />

Figure 45: Compound rims from Nusa Roviana. NR.34 has excised lines forming the<br />

triangle pattern on the upper rim. ..................................226<br />

Figure 46: Impressions on the top <strong>of</strong> the lip; and spatula impressions on the neck, thought<br />

to indicate forming <strong>of</strong> the neck using a tool. .........................227


Figure 47: Excision by rotation: one end <strong>of</strong> a small twig or rod has been poked into the<br />

clay and the other end moved in a circle, leaving a conical hole. ....... 227<br />

Figure 48: Perforation (upper hole). .................................... 227<br />

Figure 49: Examples <strong>of</strong> wavy stamping and an example <strong>of</strong> dentate-stamping. . . . 228<br />

Figure 50: Applied decoration (in combination with incised decoration) with detachment<br />

scars indicating a v-pattern, and also possible attached disc. ........... 228<br />

Figure 51: A band <strong>of</strong> fingernail impressions (opposed pinch), each <strong>of</strong> which is oriented<br />

diagonal to the CVA rather than vertically, or parallel to the CVA. ...... 230<br />

Figure 52: Deformation, perforation, and the “bnd” incomplete pattern example.<br />

........................................................... 230<br />

Figure 53: “cf” (Crow’s foot) pattern on the vessel rim above a band <strong>of</strong> pinching.<br />

........................................................... 230<br />

Figure 54: Examples <strong>of</strong> lip impression in pattern “bpi” (band parallel inner-edge <strong>of</strong> lip).<br />

........................................................... 231<br />

Figure 55: Examples <strong>of</strong> lip impressions laid out in pattern “bpo” (band parallel outer-<br />

edge). ...................................................... 231<br />

Figure 56: Fragmented examples with lip impressions assigned to “band parallel outer”<br />

pattern. ..................................................... 232<br />

Figure 57: Examples <strong>of</strong> lip impressions/incision laid out in patterns “bot” (band opposing<br />

top), “bdt” (band diagonal top) and “bpt” (band parallel top), which were regarded<br />

as equivalent in analysis due to non-exclusive nature <strong>of</strong> these descriptions.<br />

........................................................... 232<br />

Figure 58: Pattern expressed using linear arrangements <strong>of</strong> fingernail pinching. . . 233<br />

Figure 59: Miscellaneous incised patterns: MH360 is middle row, left-hand column.<br />

GW258 is bottom-right. ........................................ 233<br />

Figure 60: Examples <strong>of</strong> applied nubbins and some bounded incised patterns (MH259 is an<br />

unbounded pattern). ........................................... 234<br />

Figure 61: Decorated sherds with quartz-calcite hybrid granitic temper. ........ 234<br />

Figure 62: Unbounded incised patterns. ................................. 235<br />

Figure 63: Thin incised rims from Hoghoi. .............................. 235<br />

Figure 64: Unbounded incised pattern on the shoulder from Paniavile. ......... 235<br />

Figure 65: An example <strong>of</strong> “chv” pattern, a band <strong>of</strong> unbounded linear triangles filled by


alternating fields <strong>of</strong> parallel lines. .................................236<br />

Figure 66: Example <strong>of</strong> “chv” pattern on tall rim (pinching at neck). ............236<br />

Figure 67 Curvilinear incised patterns. ...................................237<br />

Figure 68: Crow’s foot mark, probably post-deposition scratching. ............237<br />

Figure 69: Example <strong>of</strong> cross-hatch pattern “ct3”............................237<br />

Figure 70: Cross hatch pattern “cvh”. ...................................238<br />

Figure 71: Example <strong>of</strong> pattern “gm5". ...................................238<br />

Figure 72: Sole example <strong>of</strong> pattern “rl1", a double-line arrangement <strong>of</strong> single repeated<br />

fingernail impressions. .........................................238<br />

Figure 73 Roviana vessel completeness against sampling fraction, sampling fraction<br />

against vessel representation fraction; random assignment <strong>of</strong> shreds to vessels;<br />

mean EVE <strong>of</strong> 5.78%. ..........................................260<br />

Figure 74: Roviana vessel completeness against vessel representation fraction; random<br />

assignment <strong>of</strong> sherds to vessels; mean EVE <strong>of</strong> 5.78%. .................261<br />

Figure 76: Effect <strong>of</strong> sherd thickness on sherd strength (controlled for temper variation by<br />

using placered volcanic tempered sherds only). ......................306<br />

Figure 77: Body sherd size for the various temper classes. ...................306<br />

Figure 78: Histogram <strong>of</strong> body sherd size classes by site. .....................308<br />

Figure 79: Initial classification <strong>of</strong> vessel forms. ............................321<br />

Figure 80: Examples <strong>of</strong> unrestricted Form 2 variants: Form 2a (top); Form 2b (upper<br />

middle); Form 2c (lower middle); and Form 2d (bottom). ..............325<br />

Figure 81: Additional Form 2a Sherds (top 6); the two decorated gambrelled vessels from<br />

Nusa Roviana are a short-rim variant <strong>of</strong> Form 2a; the lower sherd is transitional<br />

between Form 2a and Form 2d, being unrestricted. ...................326<br />

Figure 82: Form 3 inverted restricted vessels. .............................329<br />

Figure 83: Form 3 or Form 4 (top left) and another example <strong>of</strong> a short-rim variant <strong>of</strong><br />

Form 2a (top right); the sole example <strong>of</strong> Form 3 with loop handle(s) (2 nd to top);<br />

the only confirmed Form 4 carinated bowl, in exotic temper (2 nd to bottom) and a<br />

large base sherd or frying pan, Form 5 (bottom). .....................330<br />

Figure 84: External neck radius, all sites, size intervals 10mm.................331<br />

Figure 85: All sites, sherds >25cm 2 , neck Hcurve intervals 10mm..............332<br />

Figure 86: Form 6 “Neck Hcurve” variation; the two top sherds are too small to get a


hand into (Form 6a), while the two lower sherds have head-sized or larger orifices<br />

(Form 6b). .................................................. 334<br />

Figure 87: Relationship between rim angle and rim height for two sites, all measurements<br />

shown. ..................................................... 337<br />

Figure 88: Rim angle and rim depth, all sites, filtered so that EVE is greater than 9%, to<br />

reduce the effects <strong>of</strong> measurement error. .......................... 338<br />

Figure 89:Body sherd curvature measurements, showing spheroidal sherds (diagonal<br />

alignment) and other canister/conical forms (e.g. s-shaped pattern <strong>of</strong> Honiavasa<br />

sherd measurements). ......................................... 340<br />

Figure 90: Comparison <strong>of</strong> body sherd curvature measurements by site. ........ 342<br />

Figure 91: Robust base sherds and cannister-shaped large body sherd (the latter having<br />

exotic quartz-calcite temper). ................................... 344<br />

Figure 92: Curvature <strong>of</strong> robust body sherds (thicker than 14mm). ............. 345<br />

Figure 93: Hard interior neck pr<strong>of</strong>ile and even interior body/shoulder pr<strong>of</strong>ile (top); a<br />

s<strong>of</strong>ter neck interior pr<strong>of</strong>ile (middle) and uneven interior pr<strong>of</strong>ile (bottom). . 349<br />

Figure 94: Hard shoulder variants <strong>of</strong> Form 6 (top and middle) as distinct from the more<br />

common s<strong>of</strong>t shoulder form (bottom). ............................. 352<br />

Figure 95: Tall, fragile everted excurvate rims. ........................... 356<br />

Figure 96: Lip impression, single band on outer edge <strong>of</strong> lip, labeled by mark section: u=u-<br />

shaped, v=v-shaped, o=oblique v, w=w-shaped, s=flat-bottomed groove. . 379<br />

Figure 97: Lip impression, single band on inner edge <strong>of</strong> lip, labeled by mark section: u=u-<br />

shaped section, etc.. ........................................... 379<br />

Figure 98: Lip impression, single band on top face <strong>of</strong> lip, labeled by mark section: u=u-<br />

shaped, etc.. ................................................. 380<br />

Figure 99: Lip impression, bands on both edges <strong>of</strong> the lip. Labeled by mark section: u=u-<br />

shaped, etc.. ................................................. 380<br />

Figure 100: Calculation <strong>of</strong> lip orientation angle. .......................... 382<br />

Figure 101: Lip orientation and location <strong>of</strong> impressions..................... 383<br />

Figure 102: Rim depth by decorative class. .............................. 384<br />

Figure 103: Rim depth and rim angle <strong>of</strong> undecorated lip-rim-neck sherds. ...... 387<br />

Figure 104: Location <strong>of</strong> bands <strong>of</strong> lip impression in relation to rim form variability.<br />

........................................................... 388


Figure 105: Neck thickness comparison <strong>of</strong> Gharanga/Kopo and Miho styles/types.<br />

............................................................389<br />

Figure 106: Neck Vcurve for Gharanga/Kopo decoration, Miho Decoration, and plain<br />

rim-plain neck-plain shoulder sherds. ..............................390<br />

Figure 107: Planilateral sectioned adze from Hoghoi initial surface collection (top), plano-<br />

convex-sectioned type V adze from Zangana (middle), and planilateral adze<br />

fragment from Zangana (bottom). .................................402<br />

Figure 108: Images <strong>of</strong> the adzes shown in the preceding illustration. ...........403<br />

Figure 109: Green “type VI” or “type VIII” triangular-section adze fragment from Miho<br />

(top); butt-end <strong>of</strong> a plano-lateral-sectioned adze from Zangana South (middle) and<br />

a fragment <strong>of</strong> a trapezoidal-sectioned Green “type IV” adze from Zangana.<br />

............................................................404<br />

Figure 110: Images <strong>of</strong> adzes illustrated on previous page. ....................405<br />

Figure 111: Canarium hammerstone from H5 ceramic findspot (top) (a similar artefact<br />

was found at Hoghoi); waisted sandstone slab from Zangana (middle); and chert<br />

flakes from Hoghoi (bottom). ....................................406<br />

Figure 112: Artefacts photographed from Oka collection and reported to be from the<br />

Paniavile site: shell and stone adzes (top); stone adzes (middle) and waisted<br />

tools/weapons and a pineapple club (bottom). ........................409<br />

Figure 113: Waisted axes photographed from the Lanni collection, courtesy <strong>of</strong> the late<br />

Mr. Phillip Lanni, found in the vicinity <strong>of</strong> Gharanga Stream. ............410<br />

Figure 114: Un-ground adze preform photographed from Lanni collection courtesy <strong>of</strong> the<br />

late Mr. Phillip Lanni, reportedly found at the Gharanga site. ...........411<br />

Figure 115: Water-rounded lithic manuports, cortex complete. ................416<br />

Figure 116: Size distribution <strong>of</strong> fractured lithic manuports, either with some cortex or<br />

without. .....................................................416<br />

Figure 117: Intertidal collection units at Zangana: numbers are values in “unit” column in<br />

table “Flat.db” appended on CD. Units without numbers are those which yielded<br />

no sherds. ....................................................427<br />

Figure 118: Spatial distribution <strong>of</strong> the sherd sample at Zangana................428<br />

Figure 119: Across-shore size sorting at Zangana...........................429<br />

Figure 120: Possible linear settlement patterning in the distribution <strong>of</strong> large sherds at


Zangana south. ............................................... 431<br />

Figure 121: Initial point-provenanced collection <strong>of</strong> decorated sherds at Zangana<br />

(subsequent collection transects shown for spatial reference). .......... 432<br />

Figure 122: Lip deformation into a wave present in both Zangana-North and Zangana-<br />

South. ...................................................... 433<br />

Figure 123: Bands <strong>of</strong> punctation were restricted to Zangana-North. ........... 434<br />

Figure 124: Unbounded linear incision or necks banded with pinching at Zangana.<br />

........................................................... 435<br />

Figure 125: Distribution <strong>of</strong> temper classes at Zangana. ..................... 436<br />

Figure 126: Detail <strong>of</strong> distribution <strong>of</strong> temper classes at Zangana-North.......... 437<br />

Figure 127: Distribution <strong>of</strong> the sherd sample at Hoghoi. .................... 438<br />

Figure 128: Lack <strong>of</strong> sherd size variation at Hoghoi, except at 35-40m (n=4). .... 439<br />

Figure 129: Larger average manuport mass from 25m to 60m at Hoghoi. ....... 440<br />

Figure 130: Large unfractured stones and small fractured stones concentrated between<br />

25m and 75m at Hoghoi, with small rounded stones more widely distributed.<br />

........................................................... 440<br />

Figure 131: Size-sorting <strong>of</strong> manuport petrographic classes at Hoghoi, suggestive <strong>of</strong><br />

different size-procurement patterns by source. ...................... 441<br />

Figure 132: Distribution <strong>of</strong> potsherd temper classes at Hoghoi. .............. 442<br />

Figure 133: Distribution <strong>of</strong> the sherd sample at Honiavasa................... 444<br />

Figure 134: Effects <strong>of</strong> wave refraction (and collection intensity?) on sherd size at<br />

Honiavasa: the western margin is exposed to waves from Honiavasa channel,<br />

which expend their energy in a swash zone at about the centre <strong>of</strong> the site at low<br />

tide. ....................................................... 445<br />

Figure 135: Distribution <strong>of</strong> pottery tempers at Honiavasa.................... 445<br />

Figure 136: Co-joining sherds from deeper western margin <strong>of</strong> Honiavasa.. ..... 446<br />

Figure 137: Distribution <strong>of</strong> pottery decorative attributes at Honiavasa.......... 447<br />

Figure 138: A method <strong>of</strong> identifying sub-fossil organic inclusions? (Image supplied by<br />

Rafter Radiocarbon Lab) ....................................... 455<br />

Figure 139: Calibration <strong>of</strong> radiocarbon determination from a charcoal inclusion in a sherd<br />

from Paniavile................................................ 455<br />

Figure 140: Calibration <strong>of</strong> a radiocarbon determination from smoke-derived carbon on a


sherd from Hoghoi. ............................................456<br />

Figure 141: AMS sample taken from blackened sherd at far left, found on the surface <strong>of</strong><br />

unit 12 at Hoghoi. <strong>The</strong> sherd to the right, found in a subsurface test <strong>of</strong> unit 15 at<br />

Hoghoi, may be from the same vessel, and also has a sooted surface. .....457<br />

Figure 142: Vessel with surface sooting from Hoghoi dated by AMS radiocarbon.<br />

............................................................458<br />

Figure 143: Attributes used in seriations, groupings explained in <strong>chapter</strong> conclusions.<br />

............................................................467<br />

Figure 144: Correspondence plot (attributes) using all attributes and forms. .....469<br />

Figure 145: Correspondence plot (sites) using all attributes and forms. .........471<br />

Figure 146: Correspondence plot (attributes) excluding form-correlated attributes<br />

............................................................474<br />

Figure 147: Correspondence plot (sites) excluding form-correlated attributes. ....475<br />

Figure 148: Correspondence plot (sites), Honiavasa sample and form-correllated attributes<br />

omitted from data-set. ..........................................478


CHAPTER 1:<br />

RESEARCH QUESTIONS AND METHODOLOGY<br />

Introduction:<br />

I went to Roviana Lagoon as a student in a research team looking for Lapita pottery,<br />

among other things. We were faced with a conundrum, in that anything that looked<br />

vaguely like Lapita was in the sea. How to proceed? It became clear that this pattern,<br />

beyond being a nuisance, posed key research questions fundamental to archaeological<br />

interpretation in the region for the Lapita period.<br />

Lapita:<br />

Lapita pottery is a component <strong>of</strong> an archaeological horizon-style found from the Bismarck<br />

Archipelago to Samoa, dating to approximately 1300BC-800BC. Lapita pottery marks the<br />

first human colonization <strong>of</strong> Remote Oceania, that part <strong>of</strong> the Pacific that cannot be<br />

reached other than by making lengthy ocean crossings (Green 1991b) indicating that this<br />

pottery style was correlated in Remote Oceania with a maritime colonizing cultural<br />

adaptation and a period <strong>of</strong> rapid expansion (Green 1991a). Near-Oceania has by contrast<br />

been occupied for about 30 000 years, at least as far to the southeast as Buka, to the north<br />

<strong>of</strong> Bougainville (Allen et al. 1989, Wickler 1995, 2001, Wickler & Spriggs 1988), thus<br />

requiring a more complex model <strong>of</strong> local formation <strong>of</strong> the maritime colonizing adaptation,<br />

incorporating intrusive Island-Southeast-Asian Neolithic elements, local innovations, and<br />

integration with pre-existing indigenous populations (Green 1991a).<br />

1


Lapita and Solomon Islands: <strong>The</strong> Lapita Gap:<br />

Establishing the distribution <strong>of</strong> the Lapita Pottery horizon has been a major goal <strong>of</strong> Pacific<br />

archaeology since the 1950s. <strong>The</strong> Near Oceanic Solomon Islands (the Solomon Islands<br />

including Bougainville and excluding Ontong Java, Te Motu Province and Rennell-<br />

Bellona) (Figure 1) comprise a major and puzzling gap in the recorded distribution <strong>of</strong><br />

early-Lapita pottery sites (Green 1978, Kirch 1997:53, Kirch & Hunt 1988, Roe 1992,<br />

1993, Spriggs 1997: 128). Ultimately at issue is whether the distribution <strong>of</strong> early-Lapita<br />

pottery was continuous or discontinuous in Near Oceania. This issue has significant<br />

implications for our ideas about what Lapita represents. A continuous distribution <strong>of</strong><br />

early-Lapita pottery across this region would favour a model <strong>of</strong> the largely indigenous<br />

development <strong>of</strong> Lapita pottery, or at least raise the probability <strong>of</strong> integration into local<br />

populations <strong>of</strong> any migrants from Island Southeast Asia at an early stage in the<br />

development <strong>of</strong> the Lapita cultural complex. A discontinuous distribution would favour<br />

an avoidance model <strong>of</strong> Lapita colonization (e.g. Sergeantson & Gao 1995: 169), where<br />

Lapita represents “foreign” intrusion, by an expansionist society living at the fringes <strong>of</strong><br />

an already occupied hostile and/or malarious Near-Oceanic Solomon Islands. In this sort<br />

<strong>of</strong> model Lapita expansion “colonies” are seen as confined mostly to <strong>of</strong>fshore islands in<br />

the Bismarck Archipelago, and remain culturally and genetically relatively distinct from<br />

earlier occupants <strong>of</strong> Near Oceania for an extended period; generally bypassing the bulk<br />

<strong>of</strong> the Near Oceanic Solomon Islands in favour <strong>of</strong> a previously unoccupied or otherwise<br />

healthier Remote Oceania.<br />

2


Figure 1: Near Oceania, Remote Oceania and Solomon Islands, showing the location <strong>of</strong> the New Georgia Group.


Opinion amongst archaeologists with an interest in Lapita is divided as to the reasons for<br />

the gap in the distribution <strong>of</strong> early Lapita sites. Two possible explanations commonly<br />

discussed previously are (a) that the gap is purely an artefact <strong>of</strong> insufficient survey in this<br />

region (Green 1978, Spriggs 1997:128) and (b) that the gap directly reflects a<br />

discontinuous distribution <strong>of</strong> Lapita in this region or even complete avoidance <strong>of</strong> this<br />

region by Lapita peoples in the past (Gorecki 1992, Roe 1992, 1993:185, Sheppard et al.<br />

1999). Some take an equivocal position in relation to these possibilities (e.g. Kirch<br />

1997:53) (but see also Kirch 1997:283). An additional possibility (c), less <strong>of</strong>ten considered<br />

is that tectonic instability has reduced coastal site visibility/preservation in the Near-<br />

Oceanic Solomon Islands (Dickinson & Green 1998, Kirch & Hunt 1988:18). To these a<br />

fourth possibility is added here (d): that the primarily terrestrial archaeological survey<br />

which has occurred in the Solomon Islands has failed to locate Lapita sites due to a pattern<br />

in the past in this region <strong>of</strong> Lapita settlements located exclusively over the intertidal zone,<br />

comprising stilt houses or small artificial islets. Implicit in this proposition is the hypothesis,<br />

related to (c) above, that archaeological deposits resulting from intertidal-zone settlement<br />

in tropical waters are easily destroyed or hidden by even slight changes in sea-level, and<br />

are subject to destruction by wave action along most coastlines. A case is made for this<br />

fourth possibility in this thesis.<br />

<strong>The</strong> Lapita Ceramic Series:<br />

While Lapita as a defined stylistic horizon has its uses, a longitudinal (temporal) view <strong>of</strong><br />

Lapita as an evolving ceramic series with regional expressions is a major research objective<br />

for many. Although considerable effort has been expended to construct and compare<br />

regional chronologies for the Lapita period (Anson 1983, 1986, 2000, Best 1984, 2002,<br />

Green 1978, 1979, Sand 2000, Specht 1969, Spriggs 1990, Summerhayes 2000a, 2002,<br />

Wickler 1995, 2001) the temporal dimension <strong>of</strong> the Lapita horizon and its aftermath<br />

remains poorly defined in many areas, and this is especially true for the Western Province<br />

4


<strong>of</strong> Solomon Islands. A major focus <strong>of</strong> any work concerned with the Lapita period must<br />

therefore be understanding ceramic variability, including the definition <strong>of</strong> temporal<br />

variability. <strong>The</strong> reliability and resolution <strong>of</strong> such constructs <strong>of</strong> ceramic variability are also<br />

key issues. <strong>The</strong>se subjects are treated in more detail in Chapter 2.<br />

Archaeology <strong>of</strong> the Intertidal Zone and Subtidal Reef Flat:<br />

<strong>The</strong> first questions that tend to arise in conversations about an archaeological record<br />

written mainly from the intertidal-zone concern cultural formation processes: whether the<br />

materials in the sea got there as a result <strong>of</strong> discard from settlement over water, or whether<br />

they were re-deposited in the sea, either as a result <strong>of</strong> human agency or as a result <strong>of</strong><br />

erosion or subsidence. A further question contingent on the answer to these is whether<br />

cultural or activity patterning is retained in the spatial structure <strong>of</strong> intertidal scatters, or<br />

whether purely geomorphological processes such as sediment size-sorting are evidenced.<br />

An understanding <strong>of</strong> both cultural and natural formation processes <strong>of</strong> intertidal<br />

sites is critical to interpretation <strong>of</strong> the Lapita gap. Models formulated in this regard,<br />

especially those concerning the state <strong>of</strong> preservation <strong>of</strong> sites in relation to degree <strong>of</strong> wave<br />

exposure, have fundamental implications for all archaeologists working in Near Oceania<br />

with an interest in the Lapita period.<br />

A related question on which higher level cultural or behavioural inferences depend<br />

is “what sort <strong>of</strong> samples are obtainable from these sites?” If we wish to obtain a saturated<br />

sample <strong>of</strong> production diversity, for example (as might be used in a seriation analysis), we<br />

might want to know whether these site-samples are representative <strong>of</strong> production in the<br />

past, or whether they are subject to some systematic or predictable biases. What is the<br />

sample size in terms <strong>of</strong> vessels, as represented by sherds? Is there evidence <strong>of</strong> historical<br />

and heritable continuity linking all site samples, or are there discontinuities in the overall<br />

sample in either <strong>of</strong> these respects? How can sample size be quantified from a collection<br />

<strong>of</strong> sherds?<br />

5


It is in the nature <strong>of</strong> surface archaeology as opposed to stratigraphic excavation to<br />

go out rather than down, in search <strong>of</strong> horizontally structured variability rather than<br />

superposition-structured variability (although intertidal-zone materials are, strictly<br />

speaking, benthic-collected rather than surface-collected, the terrestrial terms surface-<br />

collection and surface archaeology will be widely used in this thesis, to avoid unnecessary<br />

jargon). <strong>The</strong> benefits <strong>of</strong> going out rather than down are manifold when the aim is to make<br />

the region the primary unit <strong>of</strong> analysis (as advocated by Binford 1964): samples from large<br />

areas are easily obtainable, in turn allowing more sites to be collected within the resources<br />

available, which in turn permits a closer approach to the ideal <strong>of</strong> a saturated sample,<br />

particularly if people in the past tended to move periodically rather than stay in one place.<br />

When doing surface archaeology in the sea, dates by stratigraphic association are<br />

not to be had, requiring direct dating <strong>of</strong> materials adhering to artefacts or incorporated<br />

within artefacts. While this has its difficulties, in other ways it is no bad thing, since dating<br />

by stratigraphic association can be problematic (Feathers 1997), particularly for small<br />

carbonized fragments potentially affected by turbation. In stratigraphic excavation there<br />

is no remedy in collecting large numbers <strong>of</strong> samples from chronostratigraphic units/strata:<br />

the problem is only worsened, with no way to know whether a date should be accepted as<br />

associated with the materials in the unit or rejected unless its age is outrageously<br />

incongruous. A number <strong>of</strong> direct dates <strong>of</strong> various sorts were obtained in the present study,<br />

which is encouraging, suggesting the prospects <strong>of</strong> surface archaeology are much better<br />

than would have been imagined a few decades ago.<br />

<strong>The</strong> Research Region:<br />

<strong>The</strong> New Georgia Group is situated in the northwest <strong>of</strong> the Near Oceanic Solomon<br />

Islands, in the Western Province. At a latitude <strong>of</strong> eight degrees south, climate at low<br />

6


altitude is torrid-equatorial in the summer months, relieved on the coast by sea-breezes<br />

during the day, while the southeast trade-wind blows during the winter months, bringing<br />

comparatively mild showery weather. It is rare for cyclones to track directly through the<br />

Western Province (Bath & Deguara 2003), but strong winds are <strong>of</strong>ten experienced during<br />

the late summer when cyclones typically pass to the south in the region <strong>of</strong> Rennel-Bellona.<br />

A fuller discussion <strong>of</strong> winds, tides and climate is given in Chapter 6.<br />

Most <strong>of</strong> the New Georgia Group is covered in lowland primary rainforest, which<br />

has been the focus <strong>of</strong> extensive log extraction in recent decades. Inland areas were<br />

apparently intensively settled in past centuries and irrigation terraces can still be seen. <strong>The</strong><br />

modern population <strong>of</strong> around 20 000 is concentrated in coastal villages <strong>of</strong> a few hundred<br />

people, and in the main towns <strong>of</strong> Gizo, Munda and Noro. Coastal areas with good<br />

gardening soils (alluvium or uplifted backreef sediments) have a mosaic <strong>of</strong> gardens and<br />

regenerating bush. Various forms <strong>of</strong> indigenous plantation and subsistence arboriculture<br />

occupy significant tracts <strong>of</strong> land near the coast. A commercial plantation timber company<br />

operates on Kolombangara. Most local transport is by outboard-powered canoe, except<br />

within local centres or in logging areas.<br />

Geology, Geomorphology and Related Human Geography:<br />

Geological exploration has been comprehensive, beginning in the 1880s, with general<br />

descriptions and with the details filled in by the Solomon Islands Geological Survey from<br />

the 1950s till independence. Interests first in bauxite prospecting and more recently in gold<br />

prospecting have seen many prospecting-licence surveys in the 1970s and 1980s.<br />

Additional field survey and a synthesis <strong>of</strong> existing information were undertaken by the<br />

British Geological Survey between 1976 and 1983 (Dunkley 1986).<br />

<strong>The</strong> New Georgia Group was formed during an episode <strong>of</strong> volcanic activity which<br />

began in the upper-Miocene and continues today (Dunkley 1986:7), and there is rapid and<br />

substantial uplift <strong>of</strong> a forearc or outer arc region comprising Ranongga, Tetepare and the<br />

7


southern part <strong>of</strong> Rendova (Figure 2), beneath which very young, warm and ductile<br />

buoyant lithosphere <strong>of</strong> the spreading Woodlark basin is being subducted. A notable feature<br />

<strong>of</strong> this process relevant to the archaeology is the marked reduction in seismicity relative<br />

to Bougainville or Guadalcanal (Dunkley 1986:9). although a serious earthquake caused<br />

shoreline changes on Vella Lavella in 1959 (Stoddart 1969b:378).<br />

<strong>The</strong> group comprises a large number <strong>of</strong> islands trending NW-SE over a distance<br />

<strong>of</strong> 230km with a total landmass <strong>of</strong> 5,060 km 2 (Dunkley 1986:abstract), in an area <strong>of</strong> land<br />

and sea totaling 15,000 km 2 . <strong>The</strong> group is within 24-hour paddling distance (50km) <strong>of</strong> the<br />

main Choiseul/Ysabel chain, 50km distant across the New Georgia Sound to the northeast.<br />

<strong>The</strong> group is exposed to the Solomon Sea and Coral Sea to the southwest, across which<br />

lie the Louisiade Archipelago and the Queensland coast, at distances <strong>of</strong> 540km and<br />

1600km respectively, on what would have been an ideal return-voyaging course in relation<br />

to the Southeast trade-wind (Irwin 1992).<br />

<strong>The</strong> main volcanic chain (Vella Lavella, Kolombangara, New Georgia, Vangunu,<br />

Nggatokae and Northern Rendova) comprises a series <strong>of</strong> remnant composite volcanic<br />

cones <strong>of</strong> basaltic composition, with various small intrusive complexes, fringed by an<br />

intricate system <strong>of</strong> uplifted Pleistocene coral reefs and lagoons (Dunkley 1986: 12). <strong>The</strong><br />

largest <strong>of</strong> the volcanic cones is Kolombangara, rising to 1760m, while others are in an<br />

advanced state <strong>of</strong> erosion, exposing central sub-volcanic intrusion complexes. Upon<br />

examining charts in 1928, W.M Davis noted that “....none <strong>of</strong> the elevated reefs <strong>of</strong> the <strong>of</strong><br />

the Solomon Islands is more remarkable than the emerged barrier reef which skirts...New<br />

Georgia” (Davis 1928:397-398), and this complex was the subject <strong>of</strong> a more detailed field<br />

study by the Royal Society Expedition to the British Solomon Islands in the mid 1960s<br />

(Stoddart 1969a). A more recent project explicates a detailed sea-level history for this<br />

formation (Mann et al. 1998).<br />

8


9<br />

Figure 2: Map <strong>of</strong> New Georgia Group showing principal geological formations (after Coulson, Dunkerly, Hughes and Ridgeway 1987).


<strong>The</strong> Roviana Formation:<br />

<strong>The</strong> archaeological materials that form the data <strong>of</strong> this thesis are all found on intertidal or<br />

shallow-water reef flats. “<strong>The</strong> differential uplift <strong>of</strong> the New Georgia barrier reefs has led<br />

to the development <strong>of</strong> a wide series <strong>of</strong> reef shores, from vertical and overhanging cliffs to<br />

horizontal intertidal flats. (Stoddart 1969b:371).” <strong>The</strong> height <strong>of</strong> the raised-reef chains<br />

around New Georgia vary along their length and with proximity to the mainland, with local<br />

variations in uplift due to tilting and displacement along faults (Dunkley 1986:36). Of the<br />

upraised reefs forming lagoons around New Georgia, only the barrier reefs <strong>of</strong> the Roviana<br />

Lagoon are suitable for gardening, with extensive areas <strong>of</strong> Pleistocene upraised backreef<br />

or lagoonal sediments (see geological map and Dunkley 1986:37). <strong>The</strong>se may have<br />

originated as pools formed by double-barrier reef systems, which are an unusual<br />

characteristic <strong>of</strong> the New Georgia barrier system, and which testify to a complex interplay<br />

between tectonic subsidence followed by uplift and fluctuating relative sea-levels (Mann<br />

et al. 1998).<br />

<strong>The</strong>re are extensive areas <strong>of</strong> uplifted backreef sediments in some parts <strong>of</strong> the New<br />

Georgia mainland also, which are elevated to a height <strong>of</strong> 190m above sea level in the<br />

Munda and Kazukuru areas, unlike the barrier-reef chain. It is more usual elsewhere<br />

around New Georgia for the tops <strong>of</strong> the upraised barrier reef to be bare <strong>of</strong> soil, with<br />

solution erosion providing a jagged micro topography, on which only trees will grow. This<br />

seems to be the case for all <strong>of</strong> the upraised barriers around Marovo, Ngerrasi and<br />

VonaVona lagoons. <strong>The</strong> raised reefs <strong>of</strong> the Roviana formation are typically indurated and<br />

recrystallized, with karstic features resulting from partial dissolution. Ages are uppermost<br />

Pliocene in some cases on fossil evidence (Dunkley 1986:38)<br />

Deltaic fans <strong>of</strong> volcanic alluvium are found at the mouths <strong>of</strong> the larger rivers that<br />

drain the New Georgia landmass, the higher well-drained central areas <strong>of</strong> which provide<br />

favoured garden areas to the modern population. <strong>The</strong> actively prograding seaward margins<br />

10


<strong>of</strong> these deltas are s<strong>of</strong>t mud, and are not used at present for agriculture, other than Sago<br />

plantation for ro<strong>of</strong>ing and walling material. Gardens are also located at the mouths <strong>of</strong><br />

smaller streams along the mainland shore, where the shore is more accessible by canoe, and<br />

where swamp-taro may be grown in pits dug inland <strong>of</strong> the shoreline.<br />

In Roviana lagoon, extensive areas <strong>of</strong> coastal saline flats with fresher uplifted<br />

sandy/coralline lagoonal sediments containing fragmented Acropora corals are cyclically<br />

hoed, and planted in sago, coconuts and bananas. <strong>The</strong>se flats are roughly 30cm above the<br />

high tide mark, and may be inundated by saltwater during storm surges or other extreme<br />

tides. <strong>The</strong>se are most likely evidence for a recent high-stand <strong>of</strong> sea-levels (late Holocene<br />

sea level changes are quite well documented by Mann et al. 1998). Pandanus grows thickly<br />

in untended areas. Calophyllum species prosper on this flat and are an important modern<br />

economic resource, while a number <strong>of</strong> tree species used in manufacture <strong>of</strong> canoe parts and<br />

fishing spears are harvested from this flat also. Modern housing is located mostly on the<br />

uplifted backreef sediments, among gardens, but some kitchens and canoe houses are<br />

located on the lower flat adjacent to canoe wharves. Crocodiles may traverse this lower<br />

flat at night in search <strong>of</strong> penned livestock or unwary dogs or small humans. As this flat was<br />

probably inundated in the pre-1000 bp period, lagoon shorelines would have been quite<br />

steep in many places compared to their present-day conformation, making canoe<br />

landing/storage difficult in many areas, which may have been an important influence on<br />

settlement type in the Lapita period.<br />

<strong>The</strong>re are many sand keys to the west <strong>of</strong> Nusa Roviana, where tectonic uplift is less<br />

than elsewhere in the Roviana lagoon, probably as a result <strong>of</strong> faulting/tilting <strong>of</strong> the Munda-<br />

Noro reefal formation, and also to the southeast <strong>of</strong> Vangunu, at the southern margin <strong>of</strong><br />

Marovo Lagoon. <strong>The</strong>se sand keys, while picturesque in a tropical travel-brochure sort <strong>of</strong><br />

way, are sometimes swept by waves during storms, and are not favoured for settlement by<br />

local people.<br />

11


Archaeological survey for the present thesis was confined to Roviana lagoon,<br />

between Munda and Kalena bay, with intensive intertidal survey restricted to the<br />

Kaliquongu region. While there is some variation in the height above sea level <strong>of</strong> the<br />

uplifted Pleistocene reef through the survey region from Nusa Roviana to Kalena Bay, the<br />

uplifted backreef sediments and lower coastal flats are common throughout the survey<br />

region, although there are localized cliffed shores also in some places. This constancy<br />

suggests a common recent relative sea-level history to the entire survey region, borne out<br />

by the research <strong>of</strong> Mann et al. (1998). Elsewhere around New Georgia, outside the survey<br />

frame <strong>of</strong> this research, recent sea-level history is more varied.<br />

Regarding long-term (post-Pliocene) differences in uplift, to the east <strong>of</strong> the survey<br />

region, between Kalena bay and Viru Harbour, the situation differs, with uplift obviously<br />

greater, and backreef/lagoon sediments extending from the reef to the mainland, with no<br />

modern lagoon remaining. To the West <strong>of</strong> the survey region, uplift is less, with the<br />

Pleistocene reef barely emerged, forming the base <strong>of</strong> the sand keys that shelter the Munda<br />

coastline. West <strong>of</strong> the Honiavasa passage the lagoon-side geomorphological record <strong>of</strong> the<br />

uplifted barrier chain is complicated by extensive road works during the Second World<br />

War (WWII), which in places substantially modified the shore pr<strong>of</strong>ile to create roads and<br />

industrial installations (Bethlehem Island and Zangana Point are examples <strong>of</strong> major<br />

earthmoving during WWII).<br />

<strong>The</strong> outer-arc region has more sedimentary rocks and low-grade metamorphic<br />

rocks than the main volcanic chain, where these are virtually absent other than as a small<br />

formation on Kolombangara. <strong>The</strong>se sedimentary rocks and tuffs are <strong>of</strong> interest due to the<br />

presence in intertidal archaeological scatters <strong>of</strong> numerous abrader fragments and adze<br />

fragments that are either sedimentary or possibly metamorphosed sedimentary materials<br />

(see Chapter 10). <strong>The</strong> outer-arc region is also <strong>of</strong> interest due to higher rates <strong>of</strong> tectonic<br />

uplift, that have outstripped the Holocene marine transgression by hundreds <strong>of</strong> metres.<br />

12


<strong>The</strong> sea-level history <strong>of</strong> the outer-arc coastlines thus contrasts markedly with that <strong>of</strong> the<br />

New Georgia uplifted fringing reef and comparison <strong>of</strong> the archaeological record there with<br />

Roviana Lagoon is desirable, although this has not yet been undertaken. <strong>The</strong> lower two-<br />

thirds <strong>of</strong> the Kiorosi section on Tetepare is <strong>of</strong> Early Pleistocene age, beginning about 1.65<br />

Ma. at the base <strong>of</strong> the exposure, and the upper third is <strong>of</strong> late-Pleistocene age, postdating<br />

0.27Ma. (Dunkley 1986:28), and has undergone exceptionally high rates <strong>of</strong> uplift in the<br />

Late Pleistocene. It is not clear to what extent remnant uplifted shorelines are exposed on<br />

the outer arc, as massive slumping has occurred in places along the exposed southern coast<br />

<strong>of</strong> Rendova/Tetepare, although Plio-Pleistocene raised fringing reefs are well preserved as<br />

a series <strong>of</strong> terraces to the southwest <strong>of</strong> the Rendova volcanic cone.<br />

Summary <strong>of</strong> Geology:<br />

In summary the geology <strong>of</strong> the New Georgia Group presents a complex tapestry <strong>of</strong> rock<br />

types, structural geology/plate tectonics and sea-level change, dominated from the<br />

archaeologist’s point <strong>of</strong> view by raised reef formations <strong>of</strong> considerable antiquity. <strong>The</strong>re are<br />

some remaining uncertainties about the timing and magnitude <strong>of</strong> Holocene changes in<br />

relative sea-level, but there is no evidence for significant change in the overall<br />

conformation <strong>of</strong> the Roviana Lagoon during the late Holocene, in terms <strong>of</strong> exposure to<br />

ocean storm waves at least. <strong>The</strong> Roviana Lagoon existed in approximately its present form<br />

since the Holocene marine transgression, but inundation <strong>of</strong> the coastal flat during a recent<br />

high-stand would have made many <strong>of</strong> the shorelines more difficult to occupy and land<br />

canoes on. <strong>The</strong>re seems to be no difference across the survey area in this respect, other<br />

than the down-faulted or down-warped sand keys west <strong>of</strong> Nusa Roviana, which would<br />

have been near-submerged during a 1m high-stand, making them even more exposed to<br />

storm surge than at present and probably undesirable for any form <strong>of</strong> settlement. Some<br />

cliffed lagoon-side coasts in the Aroroso Passage area that do not have<br />

13


Figure 3: <strong>The</strong> author providing scale for solution notches in Plio-Pleistocene limestone<br />

cliff near Saikile passage, Roviana Lagoon (photograph courtesy <strong>of</strong> Peter Sheppard)<br />

the slightly-raised shoreline flat are examples <strong>of</strong> local topography unsuitable for the growth<br />

<strong>of</strong> coral, along which the more usual series <strong>of</strong> shore terraces has not formed (Figure 3).<br />

<strong>The</strong> Roviana Early Ceramic Archaeological Record:<br />

Reeve reported four archaeological sites in the Western province, most notably the<br />

Paniavile intertidal site in Roviana Lagoon (Figure 4),<br />

“Containing a distinctive and possibly Lapita-related style <strong>of</strong> decorated<br />

pottery.... While the absence <strong>of</strong> dentate-stamping makes it impossible to<br />

consider the Paniavile material as belonging within the classic Lapita<br />

tradition, its decorative techniques and its wide range <strong>of</strong> vessel<br />

shapes...suggest that it represents a related tradition, possibly derived or<br />

descended from Lapita” (Reeve 1989)<br />

14


Reeve suggested that these ceramics closely post-date Lapita, and further suggested<br />

similarities to a number <strong>of</strong> ceramic assemblages from elsewhere in Melanesia. Reeve<br />

suggested that the Paniavile site in its design system is more similar to late Lapita material<br />

at Watom (Anson 1983, Green & Anson 1991). This suggestion is examined in the course<br />

<strong>of</strong> this thesis, and the Paniavile pottery is found to be quite dissimilar to Watom material<br />

(see Chapter 13), although his statement, “It may be that the Paniavile ceramics in some<br />

way represent a link between the late Lapita material... and the Mangaasi/Sohano ceramic<br />

traditions.” (Reeve 1989) is closer to the conclusions drawn in Chapter 13.<br />

Research Questions:<br />

Diverse specific research questions arise in relation to these materials in the sea. <strong>The</strong><br />

overall question being, “How to proceed?”, leads to a series <strong>of</strong> “How to proceed”<br />

questions and answers, rather than a Roviana prehistory.<br />

How can the formation <strong>of</strong> the Roviana intertidal-zone and reef-flat record be<br />

researched? What scale, method and intensity <strong>of</strong> survey is required to obtain a useful<br />

sample set? What cultural and post-depositional formation processes can be inferred to<br />

have produced the record? Do the Roviana materials show heritable continuity from<br />

Lapita? Were settlements over water or on land? Can we see settlements in the distribution<br />

<strong>of</strong> materials, or are artefact/manuport density distributions a product <strong>of</strong> post-depositional<br />

factors? What systematic biases are there in the record? How should the record be written<br />

(what collection/recording methods are appropriate; how much detail is needed)? How<br />

fragile are these sites ( for example, in what state <strong>of</strong> preservation are they on discovery,<br />

and what is the impact <strong>of</strong> surface collection)?<br />

How can we infer the distribution <strong>of</strong> Lapita-phase pottery use in the past in Near Oceania?<br />

What state <strong>of</strong> preservation are the Roviana materials in, and what are the taphonomic<br />

processes affecting them? Can we model preservation <strong>of</strong> similar sites elsewhere using<br />

these data?<br />

15


Figure 4: Paniavile at low tide, looking north, with several inhabited islets and the New<br />

Georgia mainland in the distance.<br />

How might a ceramic chronology be constructed? What classificatory units fit the<br />

sample and the period? How should the properties <strong>of</strong> samples be quantified? How can the<br />

various dimensions <strong>of</strong> time, style, space and function be defined and identified or controlled<br />

for? Are the available samples adequate for the task? If not, how might a better sample be<br />

obtained? What are the prospects for obtaining direct dates, and what methodological<br />

development research is needed in this respect?<br />

What can we infer about social interaction from transport <strong>of</strong> lithic raw materials and<br />

pottery temper, and how might these results be extended in future? What are the priorities<br />

for sampling <strong>of</strong> geological source rocks and sediments in future fieldwork? What are the<br />

implications <strong>of</strong> the recovered materials suite for future field collection strategies?<br />

Archaeological theory and method through which these questions could be addressed is<br />

reviewed below. <strong>The</strong> review is structured into four parts: (a) approaches to classification<br />

16


in ceramic studies generally, with some reference to Oceania; (b) a review <strong>of</strong> quantification<br />

methods and methods <strong>of</strong> sample evaluation, (c) a review <strong>of</strong> seriation method and theory;<br />

and (d) review <strong>of</strong> the links between form and function. Other topics, survey and sherd<br />

taphonomy, are reviewed in detail in their own <strong>chapter</strong>s.<br />

Fundamental to the construction <strong>of</strong> temporal variability are: survey method and<br />

extent, ceramic description, classification, quantification <strong>of</strong> class membership, exploratory<br />

analysis <strong>of</strong> variability, the disentanglement <strong>of</strong> style and function, and seriation theory and<br />

method. This is especially the case where stratigraphy and superposition do not supply<br />

clear temporal information. In Oceania we tend to dutifully search out “undisturbed”<br />

deposits and avoid “disturbed deposits” within what could be called in its more extreme<br />

manifestations a “bedded strata” theory <strong>of</strong> archaeological formation, searching for stratified<br />

sequences <strong>of</strong> (especially ceramic) change. <strong>The</strong>se sorts <strong>of</strong> temporal constructions are tested<br />

against other formation theories in Chapter 2, and in the main found to be insecure when<br />

viewed from the perspectives <strong>of</strong> contemporaneous spatial diversity, settlement pattern<br />

instability, turbation, erosion, re-deposition, postdepositional stratification and sherd<br />

taphonomy, these factors acting in various combinations.<br />

Conceptions <strong>of</strong> change and classificatory systematics <strong>of</strong> Lapita studies tend to<br />

favour, or result in, coarse periodization, the temporal resolution <strong>of</strong> which is ill-matched<br />

to some <strong>of</strong> the suggested social explanations <strong>of</strong> Lapita, leading to a proliferation <strong>of</strong><br />

speculative and untested social theory. Thus some <strong>of</strong> the temporal constructions reviewed<br />

in Chapter 2 are phrased in terms <strong>of</strong> Lapita and post-Lapita; early middle and late Lapita;<br />

Early Eastern Lapita and Late Eastern Lapita, Early motifs and Late motifs, utilitarian<br />

vessel forms and non-utilitarian. Similarly, coarse approaches to time <strong>of</strong>ten involve<br />

conception <strong>of</strong> complex variables like site occupation span in terms <strong>of</strong> atemporal states,<br />

where site assemblages are “slices in time”, readable as Lapita in its early state, or Lapita<br />

in its late state.<br />

17


Two things we do know are that a style <strong>of</strong> pottery we call Lapita was widespread<br />

in Oceania about three thousand years ago and that it is not around in the same style today.<br />

Lapita has been defined in various ways, as a culture (Bellwood 1978:244), as a cultural<br />

complex and as a ceramic series (Golson 1971, Green 1992). Regardless <strong>of</strong> which<br />

definition one chooses, the weight <strong>of</strong> evidence available to us that there was a marked<br />

similarity in material culture across occupied Oceania around the first millennium BC is<br />

overwhelming to the point where this can be considered to be a fact. We also know that<br />

many non-Lapita styles <strong>of</strong> pottery have been in use in the intervening millennia, (not<br />

worrying too much about where the cut-<strong>of</strong>f point between Lapita and non-Lapita is as this<br />

is a matter <strong>of</strong> definition: non-Lapita because these do not fit the various definitions, not<br />

because they lack heritable continuity with Lapita), and can expect, except under an<br />

extreme bedded-strata formation theory, that the archaeological record will present us with<br />

a vast array <strong>of</strong> varying mixtures <strong>of</strong> various temporal and spatial components <strong>of</strong> these styles<br />

in various stratification and superposition arrangements. To suggest on the basis <strong>of</strong> these<br />

facts that we know the direction <strong>of</strong> ceramic change is to treat the record as though the<br />

direction <strong>of</strong> change is constant. Such a doctrine <strong>of</strong> gradualism seems implicit in the<br />

conclusions <strong>of</strong> some <strong>of</strong> the studies reviewed in Chapter 2, and encourages complacence<br />

regarding the resolution with which ceramic change has thus far been defined.<br />

It seems unlikely for example, if aims are pitched at these coarse levels <strong>of</strong><br />

resolution, that the challenges posed by post-processual critiques or by evolutionary<br />

archaeology to produce historical accounts and/or to model individual and political agency<br />

can be met. We condemn ourselves to explanation only at the broadest timescale, in which<br />

the material effects <strong>of</strong> individuals and smaller groups over short timescales are invisible.<br />

For Lapita, the corpus <strong>of</strong> short-term events is more limited that in many other regions and<br />

time periods around the world: Lapita burials, for example are rare or unknown; but the<br />

stuff <strong>of</strong> this thesis, the manufacture and discard <strong>of</strong> pots, comprises short term events<br />

18


structured by the actions <strong>of</strong> individuals, and rich in stylistic data which can be used to tell<br />

time, and thus there must be information to hand in the patterning <strong>of</strong> these events across<br />

the landscape that pertains to evolutionary historicist as well as humanist orientations.<br />

This information cannot be accessed though, without a fine-grained ceramic<br />

chronology. Social explanations require a measure <strong>of</strong> social time or risk reading temporal<br />

or other sorts <strong>of</strong> variability as social, or vice versa. Ceramic descriptive and classificatory<br />

units must be formulated to capture variability that is salient to the desired level <strong>of</strong><br />

analytical resolution, and yet the fineness <strong>of</strong> one’s classification is constrained in reality by<br />

the properties <strong>of</strong> that sample in terms <strong>of</strong> historical continuity and sample size. Samples<br />

without large time-gaps or space-gaps are needed if research questions are to be pitched<br />

at a humanist or social-evolutionary level. Also, the more complete the sample, the more<br />

finely it can be split into classificatory units, which also affects the resolution <strong>of</strong> seriation<br />

chronologies. Sample evaluation is thus a crucial element in establishing the type <strong>of</strong><br />

questions that can be asked <strong>of</strong> sample variability.<br />

Seriation continues to be one <strong>of</strong> the backbones <strong>of</strong> archaeological dating in the<br />

radiocarbon era, to a greater extent than was initially expected (O'Brien & Lyman<br />

2000:226), although this is not always apparent in Pacific archaeology. Over the decades,<br />

in the work <strong>of</strong> Green, Parker, Sheppard and Green, Irwin, Specht, Wickler, Egl<strong>of</strong>f,<br />

Wahome and Best reviewed in Chapter 2, surface scatters <strong>of</strong> ceramics are valued, and<br />

attempts are made to seriate these in some cases. Seriation method and theory have<br />

undergone many reincarnations through the Culture Historical era and the period <strong>of</strong> the<br />

“New Archaeology” and behavioural archaeology, and over the last decade, in which new<br />

analytical techniques have become widely available.<br />

An emerging concern with formation theory is apparent in recent decades in world<br />

archaeology, which has complicated the interpretation <strong>of</strong> archaeological variability,<br />

particularly for ceramics, and challenged our methods <strong>of</strong> identifying temporal, stylistic and<br />

19


functional variability in the archaeological record. Regional historical examples are given<br />

<strong>of</strong> Oceanic seriations in Chapter 2, which are compared methodologically with current<br />

techniques, methods and theory.<br />

A seductive notion for those using many <strong>of</strong> the newer techniques <strong>of</strong> multivariate<br />

exploratory analysis for seriation is that the theoretical dimensions <strong>of</strong> sample-size, space,<br />

time, style and function equate to summary dimensions <strong>of</strong> variability in the data<br />

(Summerhayes 2000a, Wahome 1999). Seriation is a temporal construct, and is largely a<br />

search for temporal variability, regardless <strong>of</strong> the seriation technique used. A theme <strong>of</strong> the<br />

review in Chapter 2 is to assess the degree to which researchers have sought out temporal,<br />

as opposed to other sorts <strong>of</strong> variability, with which to construct their seriations.<br />

Furthermore, form and function are not equivalent, nor are style and decoration. <strong>The</strong>re is<br />

a tendency also for archaeologists, when faced with multiple summary dimensions <strong>of</strong><br />

variability, to ascribe each <strong>of</strong> these uncritically to one <strong>of</strong> the above. Thus in a Principle<br />

Components Analysis there is a tendency to ascribe one summary dimension to function,<br />

one to sample size, and so on, <strong>of</strong>ten with little or no explicit justification. If purely<br />

temporal variability is input into such a multivariate analysis, we will still get multiple<br />

summary dimensions <strong>of</strong> variability. Some attributes may serve to differentiate some<br />

temporal units, other stylistic variants will serve to discriminate others, as has been<br />

recognized for many decades from simpler methods <strong>of</strong> seriation.<br />

Techniques like Correspondence Analysis, which provided diagnostic statistics for<br />

the purpose, allow a clear understanding <strong>of</strong> the origins <strong>of</strong> summary components <strong>of</strong><br />

variability, to a greater degree than some <strong>of</strong> the early computer multivariate methods that<br />

used similarity matrices. <strong>The</strong> era <strong>of</strong> similarity matrix clustering approaches is largely past,<br />

as such analyses are difficult to evaluate, but it will be seen that we have a substantial<br />

residue <strong>of</strong> Lapita studies <strong>of</strong> this sort, which require careful reading.<br />

<strong>The</strong> conclusions reached in the review in Chapter 2 regarding the significance <strong>of</strong><br />

20


the main constructions <strong>of</strong> Lapita and post-Lapita variability, and regarding formation <strong>of</strong><br />

the Lapita record in general terms, can be summarized in the suggestion that Lapita<br />

chronology in Near-Oceania is not yet well understood at any but the coarsest level <strong>of</strong><br />

temporal resolution, that <strong>of</strong> the culture historical horizon, and that a major methodological<br />

and theoretical re-evaluation is required in order to achieve better resolution than this.<br />

More <strong>of</strong> the same sort <strong>of</strong> data is not enough, the type <strong>of</strong> data and what we do with it must<br />

change. I thus take issue with Spriggs’ characterization <strong>of</strong> Near-Oceanic archaeology, and<br />

Lapita archaeology in particular, as being in a data-led pioneering phase (Spriggs 1993).<br />

It will remain in that phase only if we continue to approach the record as though we are<br />

in that phase. Gosden’s call for “concerted frameworks” for analysis (Gosden 1991a), is<br />

extended here to a call for more emphasis on sampling and formation frameworks, applied<br />

not just to individual site samples, but to analysis <strong>of</strong> regional archaeological landscapes.<br />

Approaches to Ceramic Classification and Analysis:<br />

Introduction:<br />

Typological approaches to the archaeological record and the construction <strong>of</strong> chronology<br />

have their origins in nineteenth-century Europe, with developments like the Montelian<br />

synthesis <strong>of</strong> European prehistory in the 1880s, with full-fledged development <strong>of</strong> a culture<br />

historical approach and the notion <strong>of</strong> “archaeological cultures” apparent by the publication<br />

<strong>of</strong> Childe’s “Dawn <strong>of</strong> European Civilization” in 1925 (Trigger 1989:163-169). Culture-<br />

historical approaches in the Americas involved typological method that attained its<br />

archaeological zenith in the work <strong>of</strong> Willey and Phillips (1958).<br />

Rice advocates what has more recently been called a reflexive approach to<br />

classification and typology (Rice 1982:49), citing Brew’s 1949 statement “We need more<br />

rather than fewer classifications, different classifications, always new classifications, to<br />

21


meet new needs”. Rice identifies the perennial issues as<br />

• Are types real or created, (essentialist vs. materialist approaches, see Dunnell<br />

1986)<br />

• Is it better to lump or split,<br />

• What and how many attributes should be used in classifications?<br />

• What kinds <strong>of</strong> types are there (historical, analytical, cultural)<br />

• What is the difference between a grouping, a classification, and a taxonomy?<br />

Beginning with the last question, Rice suggests grouping is phenomenological, groups<br />

consist <strong>of</strong> members, while classes are defined by criteria. Groups are historical or localized<br />

events, classes are ahistorical. Where attributes are <strong>of</strong> equal weight, they create unordered<br />

classes or groups, whereas where attributes or features are differentially weighted they<br />

create hierarchical classifications, or taxonomic classifications, like the type-variety system<br />

or the Linnaean system.<br />

In the type-variety system <strong>of</strong> hierarchical taxonomy, “ware” was a broad, high-level<br />

unit <strong>of</strong> synthesis and comparison, defined, for example, by such attributes as commonalities<br />

<strong>of</strong> paste and surface treatment (Rice 1982:50). Wares were composed <strong>of</strong> groups and<br />

groups were composed <strong>of</strong> types. Types were the basic American unit <strong>of</strong> description,<br />

frequently ascribed a cultural meaning as a ceramic idea <strong>of</strong> a society. Types were formed<br />

by combining varieties, the initial sorting units.<br />

Types vs Attributes:<br />

Shepard pointed out the principal limitations <strong>of</strong> types is that these are time-sampling and<br />

space-sampling-dependent, decrying the 1950s paradigm that types are in some measure<br />

cultural entities (Shepard 1963:308). She proposed the use <strong>of</strong> technical features, such as<br />

tempering material, as providing simple criteria for delimiting types, and this influence<br />

22


can be seen strongly in Oceania in Specht’s analysis <strong>of</strong> Buka ceramic variation (Specht<br />

1969). Where materials exhibit continuous gradation, these do not form separate types, but<br />

where one class <strong>of</strong> tempering material does not grade into another, this provides a useful<br />

criterion for classification.<br />

She regarded 1950s types as excessively coarse-grained and proposed “design<br />

styles” as a finer classification, defined in terms <strong>of</strong> elements and motifs, and falling within<br />

a technological tradition, this being a cluster <strong>of</strong> associated or interdependent techniques<br />

(Shepard 1963:320). Again the influence on Specht’s approach is clear. She noted Colton’s<br />

proposition from the 1950s that a “series” consisted <strong>of</strong> a temporal sequence <strong>of</strong> types, as<br />

opposed to a “ceramic group” which was a contemporary grouping <strong>of</strong> pottery types within<br />

a site <strong>of</strong> short duration and limited area, and an “index ware” which was a group <strong>of</strong><br />

functional types peculiar to a particular prehistoric tribe.<br />

Classification and the New Archaeology:<br />

In critique <strong>of</strong> the type-variety system Dunnell wrote<br />

“If classifications <strong>of</strong> any kind are to be devices useful in constructing<br />

explanations, if they are done for something other than amusement, they<br />

must be capable <strong>of</strong> evaluation, susceptible to change. In short, they must<br />

be hypotheses about the ordering <strong>of</strong> data for a specific problem....To<br />

expect that the same set <strong>of</strong> classes defined by criteria relevant to use will<br />

prove the most useful for chronology is foolish....Brew’s admonitions for<br />

more classes, not fewer, is today just as true as when it was<br />

written....Classifications need not be taken for granted. <strong>The</strong>y must suit their<br />

problem or they are useless (Dunnell 1971).”<br />

Groupings were not seen by Dunnell as an appropriate means <strong>of</strong> constructing classes, and<br />

were regarded rather as a means <strong>of</strong> generating and testing hypotheses about classes.<br />

<strong>The</strong> “New Archaeology” was a radiocarbon-dating-era development (Trigger<br />

1989:294-295) heralded by Binford as a dramatic break from the methodologies <strong>of</strong> the<br />

past (Trigger 1989:295). <strong>The</strong> New Archaeology was marked by a shift in objectives<br />

beyond chronology <strong>of</strong> archaeological “cultures” (the latter strongly identified with ethnic<br />

23


groups by the later end <strong>of</strong> the culture-historical period). Grouping rather than classification<br />

was the preferred method <strong>of</strong> analysis <strong>of</strong> variability (Graves 1998), and variation was<br />

increasingly interpreted using anthropological models. Chronology was no longer the major<br />

goal <strong>of</strong> ceramic analysis, and there was frequently an assumption <strong>of</strong> contemporaneity for<br />

pottery variety within a settlement. Ceramic design variability could therefore inform on<br />

small-scale prehistoric organizational properties.<br />

Graves’ historical account <strong>of</strong> ceramic method and theory in the American<br />

Southwest makes some points which are relevant to the Pacific and Lapita studies during<br />

the period <strong>of</strong> the New Archaeology. Graves describes this general situation in the<br />

American southwest as atemporality to a fault. Where previously culture history had<br />

employed relatively coarse-grained temporal-typological analysis <strong>of</strong> stylistic and to some<br />

extent functional variation across regions, for the early “New Archaeologists”,<br />

understanding prehistoric society through patterning in the site came to the fore. Graves<br />

identifies this phase <strong>of</strong> archaeology in the southwest with a lack <strong>of</strong> explicit consideration<br />

<strong>of</strong> site formation processes, although substantial progress was made in this regard through<br />

the 1980s.<br />

I argue in Chapter 2 that there is an “atemporality” <strong>of</strong> this sort in some analyses<br />

<strong>of</strong> Lapita variability <strong>of</strong> that era (e.g. Green 1978), which is ultimately founded on a C14-<br />

based slice in time view <strong>of</strong> site assemblages, and that this approach to chronology has<br />

become widespread in Oceanic archaeology in subsequent decades. Unlike Specht’s<br />

approach to definition <strong>of</strong> a cultural sequence for Buka, chronology was no longer the<br />

primary objective <strong>of</strong> ceramic analysis, as chronology could be derived from other,<br />

ostensibly more scientific and objective information (radiocarbon dating). Instead, ceramic<br />

variability was used to identify activity patterning or to measure the similarity <strong>of</strong> sites to<br />

infer interaction (Green 1978) rather than to tell time. Although Green more recently<br />

returned to ceramic variability for a chronology when the accumulating 14 C evidence<br />

24


egan to look less clear-cut (Green 1991c) the extent to which the understanding <strong>of</strong><br />

ceramic variability was initially predicated on a “slice-in-time” view <strong>of</strong> sites is difficult to<br />

appreciate without careful reading <strong>of</strong> Donovan’s thesis (Donovan 1973) and Green’s<br />

working paper (Green 1978).<br />

Seriation continued in widespread use around the world through the period <strong>of</strong> the<br />

New Archaeology, but with a difference. Typological classification approaches had been<br />

largely swept away by attribute-based grouping analyses (e.g. Le Blanc 1975). Orton<br />

identifies the work <strong>of</strong> Shepard as a nodal point in this transition, where in addition to<br />

typological construction <strong>of</strong> chronology, a preoccupation with the geological origins <strong>of</strong><br />

ceramic materials emerged as an indicator <strong>of</strong> trade/exchange, and where physical properties<br />

<strong>of</strong> vessels were used to infer technological character (Orton et al. 1993:13). Shepard’s<br />

design attributes were however, far removed from the decorative attribute combination<br />

approach common in Pacific archaeology since the 1970s (Best 1984, Frost 1974, Green<br />

1978, Irwin 1972, 1985), and more akin to Mead’s structural approach (Mead 1975) in<br />

that she felt we should “insist that the meaning <strong>of</strong> design should be sought wherever<br />

possible (Shepard 1963:259)”, so Orton’s review, while pertinent to Pacific analyses <strong>of</strong><br />

technological attributes, glosses over some <strong>of</strong> the details <strong>of</strong> the transition to attribute-based<br />

analyses <strong>of</strong> design systems.<br />

Shepard attributed the notion <strong>of</strong> elements being the basic, irreducible units <strong>of</strong><br />

design to Chapman (Shepard 1963:266-267), while motifs were more varied, complex and<br />

distinctive than elements, and more appropriate units <strong>of</strong> design analysis than elements for<br />

some pottery. She noted that element/motif analysis was convenient for “strictly geometric<br />

designs”. She pointed out that element analysis might be helpful in the circumstance<br />

where such designs were in a state <strong>of</strong> extreme fragmentation, in which only a small part <strong>of</strong><br />

the design appears on the fragment, but cautioned regarding higher-level decisions<br />

about which sort <strong>of</strong> analysis was appropriate to a given set <strong>of</strong> designs, that “It would be<br />

25


unfortunate if dependence on sherds should dictate methods <strong>of</strong> analysis”<br />

This is the essence <strong>of</strong> two critiques by whole-design proponents levelled at the<br />

Mead system (Mead 1975), and also the Anson system (Anson 1983, 1986, 1987) to a<br />

lesser extent (Best 2002, Spriggs 1990), that is, that these seek to analyse the complex<br />

Lapita face designs at the level <strong>of</strong> the sherd by using systematics suited to strictly<br />

geometric designs. Shepard’s comments are particularly relevant to the transition from<br />

Lapita to non-Lapita, as this is widely thought by the whole-design proponents to involve<br />

a transition from complex pictorial representations (ill adapted to element-motif<br />

systematics) to strictly geometric patterns (to which such systematics are well adapted).<br />

<strong>The</strong> historical process by which this mismatch has come about is that Mead’s initial<br />

influential formulation was for a sample <strong>of</strong> mainly geometric decoration from Yanuca in<br />

Fiji (Mead 1975:37), while subsequent sampling <strong>of</strong> Lapita in Near-Oceania has recovered<br />

complex designs with greater frequency. <strong>The</strong> Mead system <strong>of</strong> analysis has in many cases<br />

been extended to cover new complex motifs without much change to the underlying<br />

element-motif-based systematics.<br />

Shepard’s propositions regarding ceramic attributes became mainstream by the<br />

1970s, by which time some were using the term “micro-seriation”. Le Blanc, for example,<br />

considered that types in any form limited the temporal resolution <strong>of</strong> seriation, as one could<br />

only define a limited number <strong>of</strong> types, beyond which declining sample size limited their<br />

utility (Le Blanc 1975:24). Too few types, it was argued, limit temporal resolution <strong>of</strong> the<br />

seriation. This is compounded by the difficulty <strong>of</strong> assigning many sherds to type, even<br />

though such sherds may display attributes.<br />

Element/motif attribute analyses possibly make better use <strong>of</strong> the variability present<br />

in small samples, but are subject to at least as much need for either independent or<br />

intuitive selection <strong>of</strong> attributes <strong>of</strong> chronological saliency, if analytical noise or<br />

contemporaneous functional variation is not to overwhelm the chronological signal. For<br />

26


attribute-based analyses, there has been very little general discussion in recent years <strong>of</strong> how<br />

this step <strong>of</strong> the seriation process might be achieved.<br />

In spite <strong>of</strong> the dominance <strong>of</strong> attribute-based analyses, arguments for a structural<br />

approach to attribute patterning continued to be put forward during the processual era:<br />

“in pottery studies, it is not the attributes (whether technological or<br />

decorative) that carry information, but the way these attributes are<br />

patterned on a particular vessel shape. By focusing on a potsherd rather<br />

than a whole vessel, the basic unit <strong>of</strong> behaviour (the vessel shape) and the<br />

structure <strong>of</strong> the other cultural units on it are missed. Ascertaining<br />

behavioural structure from potsherds, rather the structure on the vessel<br />

shape, and attempting to derive culturally meaningful information from<br />

them is like shredding a dictionary and trying to reconstruct its organization<br />

without reconstructing the pages.” (Arnold 1985:5).”<br />

Ceramic Attribute-combination Grouping Approaches in Oceania:<br />

<strong>The</strong> first example <strong>of</strong> a computer-aided multivariate attribute grouping approach to ceramic<br />

analysis in Oceania, marking the impact <strong>of</strong> the New Archaeology on Pacific ceramic<br />

classification was the work <strong>of</strong> Everett Frost in Fiji (Frost 1974). It is important to note that<br />

the aims in these early computer studies were largely chronological, rather than to do with<br />

interaction or within-site activity patterning. Frost used Sokal and Sneath’s numerical<br />

taxonomy approach, developed for taxonomy in biology (Sokal & Sneath 1963) to classify<br />

pottery assemblages. This has historical origins (in archaeology) in non-computer<br />

numerical grouping approaches (Rouse 1960, Spaulding 1953), and marks a fundamental<br />

departure (in terms <strong>of</strong> systematics) from previous approaches to ceramic chronological<br />

analysis in Oceanic archaeology (e.g. Specht 1969). A plethora <strong>of</strong> innovations in the use<br />

<strong>of</strong> computer algorithms to sort similarity matrices had emerged in archaeology in the 1960s<br />

and early ‘70s, (Marquardt 1978:270-273), <strong>of</strong> which Frost’s work is a fairly typical<br />

application for the times.<br />

Aspects <strong>of</strong> Frost’s approach, particularly Frost’s choice and coding for the<br />

computer <strong>of</strong> ceramic sherd attributes, rapidly became semi-standardized for non-Lapita<br />

27


ceramics. This standardizing effect can be seen in a series <strong>of</strong> analyses in the years<br />

subsequent to Frost’s Fijian thesis research. Frost’s analytical scheme was implemented by<br />

Irwin for his study <strong>of</strong> the Shortland Islands sequence, with some lumping <strong>of</strong> attributes,<br />

while Irwin’s PhD research on Mailu used a similar scheme (Irwin 1972:78-79, 1985:115-<br />

116), and others followed suit (Best 1984, Egl<strong>of</strong>f 1971).<br />

While similarity grouping methods have varied more in recent studies, the method<br />

<strong>of</strong> coding ceramic attributes and attribute combinations has in some studies maintained this<br />

phylogenetic similarity (see for example Bickler 1998, Clark 1999:54, Wahome 1999:36-<br />

46). Irwin provides a detailed explanation <strong>of</strong> how this was generally done. Each<br />

combination <strong>of</strong> decorative characteristics <strong>of</strong> multiple parts <strong>of</strong> the vessel was counted, and<br />

the frequency <strong>of</strong> these combinations by site-assemblage was counted (omitting measured<br />

form variables, which were analyzed separately as an “independent test” <strong>of</strong> the other<br />

results) (Irwin 1972:77-88). In Irwin’s case this initially resulted in 300 combinations,<br />

which were further reduced by combining some rim shape classes and excluding some data<br />

from sherds too small to yield anything other than “nugatory” data. Some decorative<br />

attributes were disregarded because they were seen to co-occur with others which were<br />

retained. <strong>The</strong> number <strong>of</strong> attribute combinations was thus reduced to 103.<br />

Variation in brokenness between assemblages is potentially problematic when<br />

using this method <strong>of</strong> ceramic description, although perhaps this is what Irwin meant by<br />

“nugatory”. Two identical sherds could potentially count as three entirely different<br />

combinations if one <strong>of</strong> the sherds is broken apart. In this sort <strong>of</strong> quantification tight<br />

control for the vessel parts represented by sherds would need to be maintained to negate<br />

this. Best achieved such control by recording different parts <strong>of</strong> the vessel as separate<br />

records in his data table (Best 1984), confining his attribute list to technological attributes<br />

rather than design motifs, but here the disadvantage is the loss <strong>of</strong> information from large<br />

sherds concerning the structure <strong>of</strong> decoration across the vessel (see also Wahome<br />

28


1999:100). Thus depending on the level <strong>of</strong> control for part representation, this approach<br />

can be a structural approach to whole vessels where attribute combinations are reminiscent<br />

<strong>of</strong> types, or a part-attribute approach, akin to Arnold’s “shredded dictionary”.<br />

Other Systems <strong>of</strong> Decorative Classification:<br />

A less subjective, but potentially less behaviourally salient approach is to avoid selection<br />

<strong>of</strong> attribute combinations altogether, and simply count the occurrence <strong>of</strong> all individual<br />

attributes (e.g. Anson 1983: motif inventory appendix) (see also Wahome 1999:101-103),<br />

rather than trying to capture patterned covariance-by-part in the descriptive scheme. This<br />

approach has the flaw that the types <strong>of</strong> attributes used are unknown. Which attributes are<br />

chronologically sensitive and which vary with other factors than time? This general<br />

approach has been applied by Wickler and Summerhayes, but in both cases some form <strong>of</strong><br />

control for vessel location was maintained: in Wickler’s case this was at the level <strong>of</strong> motif<br />

counts by location, while in Summerhayes’ case broader decorative techniques, rather than<br />

individual motifs, were coded by decorative location. <strong>The</strong> difficulty here is that the larger<br />

view <strong>of</strong> patterning <strong>of</strong> decorative attributes across the vessel is lost, but one advantage <strong>of</strong><br />

such an approach is that data can be obtained from quite small sherds (in Summerhayes’<br />

case patterning <strong>of</strong> decoration across the vessel was recorded by illustrating “minimum<br />

number <strong>of</strong> vessels” (MNV) sherd sets or families).<br />

Another potential problem with the Anson/Summerhayes motif analysis is that<br />

some motifs are identifiable from very small sherds, while others are larger. This provides<br />

an additional mechanism by which motif frequencies or presence-absence may be biased<br />

by differences in brokenness between assemblages, particularly for large incised motifs, a<br />

problem noted in relation to Reef-Santa Cruz incised Lapita motifs (Donovan 1973:15).<br />

Other approaches to descriptive classification <strong>of</strong> Lapita decoration include Mead’s<br />

formulaic system <strong>of</strong> design classification (Mead 1975) (see Specht 1977 for a critique),<br />

29


and Siorat’s structurist system, whereby variants <strong>of</strong> a design are classified by reference to<br />

clues as to their sequence <strong>of</strong> execution (Siorat 1990). <strong>The</strong>re are some differences between<br />

these broadly similar approaches in how the distribution <strong>of</strong> decoration across the vessel is<br />

coded. Mead’s “design zones” are not very specific in some cases, are defined in relation<br />

to a limited range <strong>of</strong> vessel forms and are not primary to the classification <strong>of</strong> design<br />

variability into motifs. Parker, using the Mead system, noted which locations on the<br />

vessels each motif occurred on (Parker 1981), but these records were not reported on a<br />

sherd-by-sherd basis, and frequency data <strong>of</strong> decorative locations were not given. Siorat<br />

was more specific than Mead on the sequence <strong>of</strong> application <strong>of</strong> decoration, and the<br />

location <strong>of</strong> decoration, and noted the difficulties <strong>of</strong> applying such a system to very<br />

fragmented collections.<br />

Green’s synthesis <strong>of</strong> Lapita variability (Green 1978) serves as a study in contrast<br />

to the Frost/Irwin approach to description <strong>of</strong> decorative variability, although analytical<br />

methods were very much in the Frost/Irwin attribute multivariate grouping mode, and also<br />

included Irwin-style application <strong>of</strong> Renfrew and Sterud’s close-proximity analysis. One <strong>of</strong><br />

Green’s aims was to show that there was a Lapita ceramic series divisible into a western<br />

adaptation and a derivative, but isolated, Eastern Lapita. He thus wished to identify spatial<br />

and temporal components to the Lapita horizon. His construction <strong>of</strong> a Lapita Ceramic<br />

series using the Mead system <strong>of</strong> decorative classification was made to test whether ceramic<br />

variability was consistent with his theory <strong>of</strong> an isolated Eastern Lapita region.<br />

Green’s motifs were those identified by contributors to the Mead system <strong>of</strong> design<br />

classification (Donovan 1973, Mead 1975), and were a fundamental departure in ceramic<br />

decorative description from the Irwin/Frost attribute combination approach. A<br />

consequence <strong>of</strong> the schism between the Frost/Irwin approach to ceramic description and<br />

the Mead/Donovan approach was that different periods within the Oceania regional<br />

30


chronology began to be analyzed using these separate systems <strong>of</strong> ceramic design<br />

description. Frost/Irwin attribute descriptions (with modifications) were generally applied<br />

to post-Lapita ceramics (Best 1984, Clark 1999, Egl<strong>of</strong>f 1971, Wahome 1999), while<br />

Lapita-phase ceramics were analyzed using the Mead-Donovan system, or its “splitting”<br />

derivative, the Anson system <strong>of</strong> motif inventory (Anson 1983). Another Mead/Donovan<br />

derivative, (having an extra hierarchical level <strong>of</strong> motif classification), was added by Sharp<br />

(Sharp 1988, Sharp 1991), allowing a bit more <strong>of</strong> a lumping approach, always beneficial<br />

for sample sizes, but this work was never completed.<br />

Another structurist approach to ceramic attribute description, in this case to post-<br />

Lapita decoration, for which variations in brokenness between contexts can be controlled,<br />

is seen in a transformation <strong>of</strong> the Irwin/Frost data structure (Clark 1999:65-66, Appendix<br />

2). For the purpose <strong>of</strong> his multidimensional scaling analysis (MDS) <strong>of</strong> variation, Clark re-<br />

organized the attributes “surface modification position” and “surface modification” (which<br />

had followed the Irwin/Frost/Best structure in initial recording) into three attributes,<br />

“Decoration- Lip”, “Decoration-Rim” and “Decoration-Body” (a similar data structure<br />

can be found in Shennan 1997:6). In this scheme control for brokenness can be maintained<br />

more easily than in the Frost/Irwin scheme, by coding absence <strong>of</strong> a part from a sherd into<br />

this data structure as a filter attribute, or by limiting the analysis to a subset <strong>of</strong> sherds for<br />

which part representation is either complete or equal (as is also possible, but more difficult,<br />

in the Frost/Irwin system by excluding some “nugatory” combinations). <strong>The</strong> trade<strong>of</strong>f is<br />

proliferation <strong>of</strong> data fields or attributes, as each vessel part needs its own set <strong>of</strong> attributes,<br />

which can lead to an unwieldly table unless a relational database is used for recording.<br />

A potential problem with such an approach is the saliency <strong>of</strong> the vessel part<br />

classification used. If the conception by past potters <strong>of</strong> design zones/vessel parts is <strong>of</strong> a<br />

different level <strong>of</strong> precision to that <strong>of</strong> the analyst, some analytical “noise” can result.<br />

31


Examples <strong>of</strong> this will be given in the Roviana ceramic analysis in this thesis (see Chapter<br />

9). <strong>The</strong> solution to this problem lies in conducting exploratory data analysis, seeking to<br />

understand patterning in the data, covariance between decorative attributes and location<br />

on the vessel. If the analyst’s definition <strong>of</strong> vessel parts is too finely split, there should be<br />

a tendency for some decorative attributes to occur on adjacent vessel parts, and these part-<br />

attribute combinations can be lumped to boost sample size. Iterative re-classification <strong>of</strong><br />

both decoration and vessel parts is needed to capture structure in the data, <strong>of</strong> the layout<br />

<strong>of</strong> decoration across the vessel.<br />

It can be seen above that decorative classification for different periods <strong>of</strong> the<br />

Oceanic sequence has developed along different lines. Lapita and post-Lapita/non-Lapita<br />

studies have diverged in their units <strong>of</strong> description/classification, more so than in methods<br />

<strong>of</strong> analysis. One reason for this is the reduction in complexity <strong>of</strong> decorative technique and<br />

decorative patterning from Lapita to non-Lapita. Lapita technique comprises a set <strong>of</strong><br />

decorative tools; roulettes, dentate-stamps, circular stamps and carving blades. Post-<br />

Lapita, (or perhaps it would be better to say non-Lapita,) these specialist tools are absent,<br />

leaving widely available general tools such as fingernails, shell edges, sticks or leaf midribs,<br />

and simple blades or fingernails for incision (comb incision and carved paddles are<br />

exceptions to this trend). <strong>The</strong> potters’ construction tools may not have changed much<br />

(paddle and anvil technique seems to be ubiquitous in all periods, if predominant post<br />

Lapita), but the decorative toolkit became less specialized. In addition to this change to the<br />

decorative toolkit, the complex pictorial and geometric designs <strong>of</strong> the Lapita phase give<br />

way to simpler designs <strong>of</strong> a more geometric type, some <strong>of</strong> which may be found anywhere<br />

from the European Neolithic (e.g. Beaker pottery) to the late-prehistoric pottery <strong>of</strong> the<br />

Amazon Basin (e.g. Arqueologicas 1970). It is also suggested above that many analytical<br />

schemes are sensitive to differences in the level <strong>of</strong> brokenness or fragmentation <strong>of</strong> pottery<br />

assemblages.<br />

32


How then might we classify Lapita and post-Lapita pottery from Oceania to avoid<br />

making a leap in systematics in mid-analysis? What is needed is a system <strong>of</strong> design<br />

classification sensitive to temporal variability, but insensitive to variation in the level <strong>of</strong><br />

fragmentation or brokenness, and salient to the Lapita/post-Lapita transition. An example<br />

<strong>of</strong> such a design-classification approach can be seen in Wickler’s analytical classification<br />

<strong>of</strong> incised decoration into bounded incised versus unbounded incised (using a modification<br />

<strong>of</strong> Mead’s general zone marker concept), although Wickler does not state the advantages<br />

<strong>of</strong> his classification. By making a decision that this simple distinction was salient to his<br />

chronological objectives, he subsumed a range <strong>of</strong> designs within a simple binary<br />

categorization, where Mead, faced with the same sherds, would, in attempting to describe<br />

how each design was put together, have come up with a whole list <strong>of</strong> motifs and their<br />

all<strong>of</strong>orms (elaborations <strong>of</strong> the same underlying motif structure), and Anson would have<br />

been in danger <strong>of</strong> splitting even further, classifying at the level <strong>of</strong> these variants.<br />

Latitudinal bands <strong>of</strong> decoration (general zone markers in Mead’s terminology) are<br />

usually located near corner points, like carinations, necks and lips. Typically carinations<br />

and necks preserve well due to their form strength, and even in very broken assemblages<br />

one can gain a good idea <strong>of</strong> the relative frequencies <strong>of</strong> these categories. Latitudinal general<br />

zone markers are therefore identifiable to type and location on the vessel over a range <strong>of</strong><br />

brokenness in some cases, since lips, necks and carinations are identifiable even when<br />

sherds are quite small. Where larger sherds suggest these bands are continuous around the<br />

latitude <strong>of</strong> the vessel and comprising repeats <strong>of</strong> closely spaced elements, there is no danger<br />

<strong>of</strong> their identity changing with changes in brokenness, provided the spacing <strong>of</strong> elements<br />

is substantially closer than the minimum dimension <strong>of</strong> sherds. For more widely spaced<br />

bands <strong>of</strong> elements, such as when an element is only repeated three or four times around<br />

the band <strong>of</strong> latitude, a high level <strong>of</strong> brokenness (small sherd size) would result in a<br />

mixture <strong>of</strong> plain and decorated sherds from what was originally a single decorated vessel,<br />

33


so the spacing <strong>of</strong> elements in relation to sherd size must be borne in mind when using<br />

banded decoration in this manner, and the abundance <strong>of</strong> plain lip sherds would be<br />

misleading.<br />

Quantification:<br />

If similarity between pottery samples is quantified using counts <strong>of</strong> the abundance <strong>of</strong><br />

attributes, what measures <strong>of</strong> quantity are appropriate to one’s particular samples? This<br />

subject has been thoroughly reviewed elsewhere (Orton 1982, 1993, Orton & Tyers 1991),<br />

but the last word on the subject has yet to be written, and this fundamental problem can<br />

only be partially avoided by resorting to occurrence seriation (discussed below), as sample<br />

richness is related to sample size for unsaturated samples, and sample size is a<br />

quantification problem.<br />

Units <strong>of</strong> Quantification and Sample Brokenness:<br />

Orton (1993) identifies one <strong>of</strong> the principal issues in pottery quantification as whether the<br />

measure used will bias the relative abundance <strong>of</strong> types/attributes. A complicating factor<br />

is that the attributes chosen for study themselves affect which measures will bias the<br />

relative abundance. A good example <strong>of</strong> this is the relative abundance <strong>of</strong> plain versus<br />

decorated pottery. If we have a hypothetical assemblage <strong>of</strong> ten globular pots, <strong>of</strong> which five<br />

vessels each have a different band <strong>of</strong> decoration somewhere on the vessel, while five pots<br />

are plain, quantification <strong>of</strong> the attributes is simple, until we smash the vessels. <strong>The</strong><br />

complete vessels can be counted if unbroken, yielding 50% decorated. Unfortunately, if<br />

using sherd count as a measure <strong>of</strong> quantity <strong>of</strong> the attributes decorated and plain, the more<br />

we smash them, the plainer the assemblage gets, because the undecorated areas <strong>of</strong> the<br />

decorated pots are counted as plain observations. Clearly simple sherd count is the wrong<br />

34


measure <strong>of</strong> quantity for these particular attributes.<br />

If the smashed assemblage were quantified as the logical minimum number <strong>of</strong><br />

vessels (MNV) represented, the five different bands <strong>of</strong> decoration would result in a count<br />

from the pile <strong>of</strong> sherds <strong>of</strong> at least five decorated vessels, while sherds from the five<br />

undecorated vessels would be counted as a single vessel, generating an assemblage <strong>of</strong> 83%<br />

decorated pots. Clearly this would be a complete mismatch between decorative attributes<br />

and unit <strong>of</strong> quantification. <strong>The</strong>re are measures that would work for this hypothetical<br />

example, for example, the sherds could be quantified using sherd area, if we had some prior<br />

knowledge <strong>of</strong> the extent <strong>of</strong> decoration on vessels and the surface area <strong>of</strong> vessels, we could<br />

work out that each <strong>of</strong> the summed areas <strong>of</strong> the five decorative types corresponded to a<br />

single vessel, and that there was enough sherdage in total for ten vessels, thus by<br />

subtraction we might arrive at the correct answer, 50% <strong>of</strong> vessels decorated. Assuming,<br />

that is, that the recovered sample includes all sherds from each pot.<br />

<strong>The</strong> difficulty and need for prior knowledge could be substantially alleviated<br />

though, by forgetting about the relative abundance <strong>of</strong> undecorated sherds, and<br />

concentrating on the relative abundances <strong>of</strong> the five different types <strong>of</strong> decoration. If the<br />

hypothetical situation is complicated slightly though, by having one <strong>of</strong> the decorative types<br />

extend over a greater portion <strong>of</strong> the vessel than another decorative type, then the MNV<br />

measure <strong>of</strong> relative abundance will be better than the sherd count. Unfortunately, if there<br />

were five pots <strong>of</strong> one decorative type and one <strong>of</strong> each <strong>of</strong> the other four types, this would<br />

not work. In this circumstance sherd count might be giving a better approximation <strong>of</strong> the<br />

relative abundance <strong>of</strong> the vessel decorative types. <strong>The</strong> point is that quantification is a tricky<br />

business, even if one is not so naive as to wish to seriate using the relative abundance <strong>of</strong><br />

plain and decorated pottery, quantified by sherd count.<br />

Units <strong>of</strong> Quantification and Vessel Completeness:<br />

35


In the critical review <strong>of</strong> the Lapita temporal constructs in Chapter 2, quantification is an<br />

issue that arises continually, both in relation to relative abundances <strong>of</strong> decorative attributes,<br />

and in relation to quantification <strong>of</strong> sample size. Sample size(n) in the statistical sense refers<br />

to the number <strong>of</strong> observations comprising the sample. For pottery we must ask,<br />

observations <strong>of</strong> what? Of pots or <strong>of</strong> pieces <strong>of</strong> pots? <strong>The</strong> difficulty with pieces <strong>of</strong> pots is<br />

that the observations are clearly not independent if decoration occurs in repeated elements<br />

or motifs across the pot. This is why sherd count is a biased measure if the extent <strong>of</strong><br />

decoration across the pot is uneven, and why MNV is not if there is only one pot per<br />

decorative type. Pots themselves may not be independent observations <strong>of</strong> behavioural<br />

variability if production is highly standardized, thus if one potter turns out fifty identical<br />

pots <strong>of</strong> one type, while another turns out five identical pots <strong>of</strong> another, and we are<br />

interested in the question <strong>of</strong> how variable pottery production was between producers, the<br />

number <strong>of</strong> observations in the sample is two, not fifty-five.<br />

In the review that follows I try to make the point that occurrence seriations <strong>of</strong><br />

broken pottery assemblages are mis-applications <strong>of</strong> a technique, because sample size is so<br />

difficult to quantify, and yet sample richness is so <strong>of</strong>ten related to sample size. How can<br />

we tell which decorative attributes are common in some sites, and thus robust indicators<br />

<strong>of</strong> presence or absence, and which are not? Can this be done by counting the pieces or by<br />

some other measure such as MNV? I have shown above that both <strong>of</strong> these measures can<br />

be wildly misleading in some circumstances. A more fundamental question that might in<br />

many circumstances allow an assessment <strong>of</strong> sample size, is to ask “what sort <strong>of</strong> sample is<br />

this?” Can the sherds in the sample be refitted to form complete vessels, or do none <strong>of</strong><br />

them seem to match? If they don’t match, is this because edges are rounded from some<br />

taphonomic process, or is it because the adjacent sherds are absent from the sample? Even<br />

if none <strong>of</strong> the sherds has matching edges, the question <strong>of</strong> whether the sherds are likely to<br />

be from the same pot or not can still be asked. Thus one <strong>of</strong> the primary questions in<br />

36


ceramic sample evaluation is “what is the vessel completeness <strong>of</strong> the sample?”<br />

Analysis <strong>of</strong> vessel completeness can give an indication <strong>of</strong> which measures <strong>of</strong> sample<br />

size and/or relative abundance are appropriate to the specific sample. At one extreme, if<br />

there are indications that most sherds are singletons, i.e. other sherds from the same pot<br />

are not present, then sherd count and vessel count are the same, sherd count observations<br />

are likely to be independent in most cases, and can be used as a unit <strong>of</strong> measurement for<br />

relative abundances and sample size. At the other extreme, where vessel completeness is<br />

high for all vessels in the sample, a vessel-based count might be the most appropriate<br />

measure <strong>of</strong> relative abundances, and would certainly be a much better indication <strong>of</strong> sample<br />

size for occurrence seriation than sherd count. Vessel completeness information is<br />

therefore an important tool in choosing an appropriate method <strong>of</strong> sample size<br />

quantification, although vessel completeness is not always easy or even possible to<br />

measure. Where it is impossible to measure vessel completeness, sample size is difficult to<br />

quantify other than by resort to sum <strong>of</strong> some measure <strong>of</strong> estimated vessel equivalent, to<br />

calculate a minimum number <strong>of</strong> vessels.<br />

Sherd Assemblages as Samples:<br />

<strong>The</strong> development <strong>of</strong> relatively labour-intensive sherds-as-vessels approaches to<br />

quantification is in part a recognition <strong>of</strong> the difficulty <strong>of</strong> comparing assemblage<br />

compositions, unless they happen to have the same levels <strong>of</strong> completeness and brokenness<br />

(Orton 1993: 169-170), and vessel completeness cannot be measured other than by the<br />

grouping <strong>of</strong> fragments into vessel classes. If the measured or estimated levels <strong>of</strong><br />

brokenness and/or completeness differ between assemblages, we must conclude that the<br />

assemblages are not equivalent samples <strong>of</strong> the material discarded by people in the past.<br />

How numerically large is our sample <strong>of</strong> potting behaviour <strong>of</strong> the past? Can the<br />

sample be read directly as representing people's potting behaviour, or is it a sample <strong>of</strong> a<br />

37


sample <strong>of</strong> a sample, subject to a number <strong>of</strong> transformations in composition? In either case,<br />

what sort <strong>of</strong> sample (representative or biased)? What is the sampling fraction? (Again,<br />

implicit in this question is the problem <strong>of</strong> how and what we are counting).<br />

Several Pacific archaeologists have touched on these subjects over the years.<br />

Specht provided a cogent discussion (Specht 1969: 70-73), Egl<strong>of</strong>f published on the subject<br />

(Egl<strong>of</strong>f 1973) independently inventing a new measure <strong>of</strong> quantity that has subsequently<br />

been widely used. Parker was careful to provide at least two different measures <strong>of</strong> sample<br />

size in her analysis <strong>of</strong> Reef-Santa-Cruz pottery (Parker 1981). Summerhayes has given the<br />

matter some thought, particularly in regard to the effect <strong>of</strong> variation in the extent <strong>of</strong><br />

decoration across the pot on the ratio <strong>of</strong> plain to decorated sherd count (Summerhayes<br />

2000c, 2002), and Clark has noted some <strong>of</strong> the biases inherent in the MNV count (Clark<br />

1999). On the world stage, the subject has been thoroughly reviewed for pottery in general<br />

(Orton 1993), amid some pessimism that we will ever posses absolute measures or<br />

estimates <strong>of</strong> vessel populations (see for example Orton 2000).<br />

<strong>The</strong> analysis <strong>of</strong> brokenness and completeness in Chapter 5 derives from an<br />

unpublished manuscript (Felgate & Bickler n.d), a methodological paper on estimating<br />

parent vessel populations from sherd samples. This approach is a fundamental departure<br />

(enabled in part by properties <strong>of</strong> samples obtained from large-area surface collections) from<br />

the approaches usually taken in Pacific archaeology, which implicitly regard the recovered<br />

archaeological sample as the target population, by regarding the measured properties <strong>of</strong><br />

the sample as representative in an unspecified way <strong>of</strong> the properties <strong>of</strong> people’s lives in the<br />

past. We need to be more specific about the nature <strong>of</strong> that representation if our behavioural<br />

inferences based on sherd assemblages are to be robust.<br />

An estimate <strong>of</strong> the total quantity <strong>of</strong> pottery represented by the recovered sample<br />

from the Roviana intertidal-zone sites is useful firstly as it gives some idea <strong>of</strong> the potential<br />

for bias in the sample. Seriation using decorative attribute abundances will be affected if<br />

38


the relative abundance <strong>of</strong> an attribute is linked to the degree and type <strong>of</strong> postdepositional<br />

sampling. Comparison <strong>of</strong> estimated breakage population and sample assemblage size will<br />

allow some idea <strong>of</strong> sampling fraction (both in terms <strong>of</strong> the bulk <strong>of</strong> sherdage and the<br />

number <strong>of</strong> vessels), and the potential for bias. This subject will be explored further in<br />

Chapter 7, where differences in part-representation for different vessel styles will be<br />

investigated.<br />

Accumulations research (estimating the intensity/duration <strong>of</strong> occupation through<br />

the quantity <strong>of</strong> pottery in the breakage population) is possible where an estimate <strong>of</strong> the<br />

breakage population is available (see Varien & Mills 1997, Varien & Potter 1997). If an<br />

independent chronology is available that allows total duration and historical continuity <strong>of</strong><br />

occupation to be modeled, then intensity <strong>of</strong> occupation can be the focus <strong>of</strong> study. <strong>The</strong><br />

results <strong>of</strong> the analysis in this <strong>chapter</strong> will be examined from this perspective in Chapter<br />

13.<br />

Knowing how much is missing from the sample informs indirectly on<br />

postdepositional formation processes. In this thesis evidence is presented that suggests the<br />

vast majority <strong>of</strong> the sherdage discarded at the Roviana intertidal-zone sites is no longer<br />

extant or recognizable as potsherds. <strong>The</strong> extent <strong>of</strong> this loss <strong>of</strong> material is estimated and<br />

explanations for this loss are explored in Chapter 7.<br />

Interpretation <strong>of</strong> survey results requires an assessment <strong>of</strong> the probability <strong>of</strong><br />

detection <strong>of</strong> evidence: when is absence <strong>of</strong> evidence acceptable evidence <strong>of</strong> absence? <strong>The</strong><br />

sherd sampling fraction and the vessel sampling fraction are also ways <strong>of</strong> characterizing<br />

the state <strong>of</strong> preservation <strong>of</strong> the extant deposit, which says something about site<br />

detectability. To what extent are they preserved in particular circumstances? Under what<br />

general conditions <strong>of</strong> wave exposure etc. do we see these sites preserved? Developing<br />

answers to these questions is crucial to interpreting the archaeological record <strong>of</strong> the wider<br />

region. If almost all <strong>of</strong> the breakage population has vanished even in the sheltered setting<br />

39


<strong>of</strong> the Roviana Lagoon, this has fundamental implications for the interpretation <strong>of</strong> the<br />

archaeological record elsewhere in Near Oceania. Just how fragile is this site type?<br />

Sample Formation <strong>The</strong>ory : the Breakage Population:<br />

<strong>The</strong> central theoretical concept put forward (following Felgate & Bickler n.d) to provide<br />

a definition <strong>of</strong> a target population is the breakage population, a theoretical stage in the<br />

formation <strong>of</strong> an archaeological sample where vessels are broken in some unspecified<br />

manner, but are also complete, i.e. all the fragments are present (Felgate & Bickler n.d).<br />

While it is clear that an archaeological sample <strong>of</strong> vessel sherds is unlikely to have ever been<br />

all in such a state at the same time, this does not negate the utility <strong>of</strong> the concept as a<br />

target population for estimation procedures (Figure 5) (see also De Boer 1983). Points<br />

to note are that the “behavioural assemblage” is regarded as equivalent or closely related<br />

to Schiffer's “systemic context” (Schiffer 1995a) rather than being a life assemblage in the<br />

palaeoecological sense <strong>of</strong> a fossilized tableau <strong>of</strong> organisms in life position (Brenchley &<br />

Harper 1998:69), such as might be seen in parts <strong>of</strong> Pompeii.<br />

Felgate and Bickler introduced the term “breakage population” as roughly<br />

analogous to DeBoer's use <strong>of</strong> the palaeoecological term “death assemblage” and draw a<br />

distinction between the breakage population and DeBoer's “discard assemblage” or<br />

Schiffer’s “discard flow”. <strong>The</strong> breakage population is defined as a theoretical time-<br />

accumulated set <strong>of</strong> the broken, but complete, pottery vessels that form the parent vessel<br />

population <strong>of</strong> a sample <strong>of</strong> potsherds from an archaeological context. This definition<br />

assumes that the use-life <strong>of</strong> a vessel ends upon breakage, but is unaffected by re-use <strong>of</strong><br />

some portions <strong>of</strong> the vessel as a secondary artefact or recycling into grog temper. DeBoer<br />

(1983) sometimes refers to a “total discard assemblage” which would seem to be<br />

40


effectively equivalent to the breakage population.<br />

Use-life seems to be the most important factor linking composition <strong>of</strong> behavioural<br />

assemblages with breakage populations (David & Wilson 1999, de Barros 1982, Mills<br />

1989, Shott 1989b, 1996). <strong>The</strong> relationship between behavioural assemblages and discard<br />

assemblages is more complex than that between behavioural assemblages and breakage<br />

populations, involving a number <strong>of</strong> factors in addition to use life, such as discard practices<br />

and re-use <strong>of</strong> fragments. By separating breakage and discard, a parent population is<br />

modeled in the formation <strong>of</strong> the archaeological record in which artefacts can exist beyond<br />

their use-life, but in a complete state.<br />

<strong>The</strong> composition <strong>of</strong> this breakage population has been structured by use-life, but<br />

with vessel completeness at 100%, unaffected by any re-use <strong>of</strong> sherds, etc. No statement<br />

regarding the degree <strong>of</strong> brokenness <strong>of</strong> this population is implied here beyond the<br />

expectation that the vessel is cracked beyond use or broken into any number <strong>of</strong> sherds.<br />

Comparison <strong>of</strong> average vessel completeness in the archaeologist’s sample with average<br />

vessel completeness in the breakage population (the latter is by definition 100%) is the key<br />

to estimating the breakage population.<br />

It is worth reiterating that the target population in the estimation approach <strong>of</strong><br />

Felgate and Bickler is the breakage population, rather than any <strong>of</strong> the other “stages” in the<br />

model <strong>of</strong> formation given in Figure 5. <strong>The</strong> estimate obtained does not inform directly on<br />

the extent <strong>of</strong> the archaeologist's sampling <strong>of</strong> the extant deposit, or on taphonomic<br />

transforms, nor does such an estimate provide information on use-life factors as these<br />

transform a behavioural assemblage into a breakage population.<br />

To show how vessel completeness can be used to estimate a breakage population,<br />

41


Felgate and Bickler turn to an explanation <strong>of</strong> similar estimation in faunal analysis. Orton<br />

explains:<br />

“<strong>The</strong> formula was based on the idea <strong>of</strong> matching pairs <strong>of</strong> bone<br />

elements; in a population, there are equal numbers <strong>of</strong> left and right<br />

elements. As the population is reduced by sampling effects, some<br />

left and some right elements are removed, leaving unmatched left<br />

elements, unmatched right elements, and matching pairs. <strong>The</strong><br />

smaller the sampling fraction, the smaller the proportion <strong>of</strong><br />

matching pairs is likely to be. <strong>The</strong> numbers <strong>of</strong> unmatched left and<br />

right elements, and <strong>of</strong> matching pairs, can thus be used to estimate<br />

the original size <strong>of</strong> the population, given assumptions about the<br />

nature <strong>of</strong> the taphonomic process (Krantz 1968:215). (Orton<br />

2000:54).<br />

Pots do not break naturally into matching pairs, unlike skeletons, but the model <strong>of</strong> sample<br />

formation described by Orton can be expanded to any archaeological fragmented materials<br />

in which essential classes exist, like collections <strong>of</strong> potsherds, in which essential vessel<br />

classes must exist (even though these may be difficult or impossible to identify in many<br />

cases). In general terms, the model <strong>of</strong> sample formation described by Orton can be<br />

rephrased as the problem <strong>of</strong> estimating the number <strong>of</strong> species or classes in a population,<br />

based on the properties <strong>of</strong> the sample.<br />

This is not just an archaeological problem; nonparametric statistical methods exist<br />

for solutions applied to a wide range <strong>of</strong> such estimation problems. Felgate and Bickler refer<br />

to the vast literature on capture-recapture studies where capturing and recapturing <strong>of</strong><br />

tagged animals can be used to estimate populations <strong>of</strong> animal species in a landscape (see<br />

Seber 1986 for a review). A review <strong>of</strong> the related problem <strong>of</strong> the estimation <strong>of</strong> number <strong>of</strong><br />

species/classes in a population (Bunge & Fitzpatrick 1993) noted a range <strong>of</strong> working<br />

solutions to the “number <strong>of</strong> species” problem, and noted prospects for improvements in<br />

this regard in the future (see also Solow 1994 for a Bayesian approach).<br />

<strong>The</strong> question <strong>of</strong> the types <strong>of</strong> bias likely to be present in samples becomes important<br />

in the context <strong>of</strong> samples comprising mostly singleton sherds. This is one <strong>of</strong> the questions<br />

42


investigated in Chapter 7. <strong>The</strong> other major question raised by the analysis in Chapter 5,<br />

also addressed in Chapter 7, is, “What happened to the pottery that didn’t make it into our<br />

sample?”<br />

For seriation, if it can be shown through an analysis <strong>of</strong> vessel completeness that<br />

most decorated sherds are likely to be singletons, or independent observations <strong>of</strong> the nature<br />

<strong>of</strong> vessels in the breakage population, then sherd count is a useful measure <strong>of</strong> sample size<br />

and relative abundance <strong>of</strong> attributes. An argument will be put forward in Chapter 5 to<br />

suggest that this is the case for the Roviana early ceramic samples from the sea. <strong>The</strong> same<br />

argument is applicable to the calculation <strong>of</strong> sample sizes. If decorated sherds are mostly<br />

independent observation <strong>of</strong> vessel properties, then sherd counts <strong>of</strong> decorative and form<br />

attributes are good indications <strong>of</strong> the number <strong>of</strong> vessels represented in the samples. This<br />

is important for seriation analysis, in which, as will be detailed in the following section,<br />

attribute sample sizes weight the analysis and affect the outcome.<br />

Seriation <strong>The</strong>ory : a Review:<br />

Seriation is a term in use principally in the fields <strong>of</strong> psychology and archaeology. In<br />

psychology it refers to the cognitive ability to order objects based on similarity, but in<br />

archaeology the meaning is more specific. When archaeologists talk about seriation they<br />

usually mean (or should mean) the creation <strong>of</strong> a temporal series based on similarity and<br />

sometimes evolutionary hypotheses.<br />

Types <strong>of</strong> Seriation:<br />

A recent taxonomy <strong>of</strong> seriation techniques divides seriation into similarity approaches, and<br />

evolutionary approaches, the latter based on a rule <strong>of</strong> evolutionary development.<br />

43


Figure 5: Processes <strong>of</strong> formation <strong>of</strong> the Archaeological record (After<br />

Felgate and Bickler n.d., adapted from De Boer 1983): inferring a<br />

breakage population from an archaeological sample.<br />

Similarity approaches are further subdivided into frequency seriation, occurrence seriation<br />

and phyletic seriation (O'Brien & Lyman 2000:64). All <strong>of</strong> these approaches have been<br />

applied to Lapita, in various mixtures, with little explicit awareness or discussion <strong>of</strong> the<br />

differences between them.<br />

Evolutionary seriation is one in which a constant direction <strong>of</strong> change or an<br />

44


evolutionary rule <strong>of</strong> development is specified. Summerhayes’ statement, for example, that<br />

his stratigraphic data identifies the direction <strong>of</strong> change (Summerhayes 2000a:151) has an<br />

implicit evolutionary component.<br />

Seriation involves firstly description <strong>of</strong> ceramic assemblages in terms <strong>of</strong> criteria<br />

(types or attributes) that are thought to vary over time (Cowgill 1972, Duff 1996, Dunnell<br />

1970, Lyman et al. 1998, Marquardt 1978, O'Brien & Lyman 2000, Shennan 1997:342).<br />

If a randomly selected set <strong>of</strong> descriptive attributes is used rather than a set selected as<br />

representing temporal change, an “...inseparable hodgepodge...” <strong>of</strong> geographic, functional<br />

and temporal variation might result (Dunnell 1970:310). (See also Braun 1983:113 on the<br />

confusion <strong>of</strong> functional and decorative variation, and the additional factor <strong>of</strong> the<br />

relationship between vessel size and decoration). While form/function/technological<br />

attributes may vary over time (Rice 1982:48), these should only be used in seriation if<br />

independent evidence is available to show that this was the case. Also, pottery production<br />

styles can be expected to vary across space. Braun also notes the need, given different use<br />

and breakage rates in different contexts, to seriate comparable contexts <strong>of</strong> deposition, and<br />

the need to control for vessel part in treating decorative variability (Braun 1983).<br />

Approaches to the identification <strong>of</strong> chronological variability range from intuitive<br />

examination in combination sometimes with the direct historical method <strong>of</strong> working back<br />

from the recent past as in Kroeber’s original formulation (O'Brien & Lyman 2000:111-<br />

114), to sophisticated computer techniques <strong>of</strong> data exploration. A simple method,<br />

sometimes used by archaeologists in the culture historical era, was to control for functional<br />

variation by performing seriation on a single vessel form, and to control for spatial<br />

variation by limiting the seriation to sites from a small region, regarding the remaining<br />

variation as temporal variation in style.<br />

Various people have pointed out that seriation is best carried out on sites that<br />

represent a short occupation, or on sites with similar occupation spans (O'Brien & Lyman<br />

2000:117). Marquardt (1978) takes a more relaxed stance on this issue than Dunnell<br />

45


(1970). Ideally one would like large samples from short, well-preserved occupations, and<br />

these may well be conditions which are met in the archaeological record in some parts <strong>of</strong><br />

the world. This has yet to be securely demonstrated in the Oceanic Lapita examples<br />

reviewed in Chapter 2.<br />

In traditional graphical frequency seriation, assemblages are graphed as relative<br />

frequencies summing to 100%. Bars representing the relative proportions <strong>of</strong> selected types<br />

or attributes are sorted until unimodal “battleship curves” are produced, usually interpreted<br />

as representing the waxing and waning popularity through time <strong>of</strong> items meeting the<br />

classification criteria. (Marquardt 1978:261). This “popularity principle” is <strong>of</strong>ten<br />

incorrectly stated as an assumption <strong>of</strong> the technique. Other interpretations are possible,<br />

relating to deposition similarity rather than directly to production similarity and popularity<br />

(Goldmann 1971). Le Blanc notes in regard to this “popularity principle”:<br />

“Even if... changes were for all practical purposes instantaneous, the<br />

archaeological record <strong>of</strong> these changes will almost invariably show them as<br />

gradual with periods <strong>of</strong> mixed technology or style (Le Blanc 1975:23).”<br />

McNutt shows that while temporal ordering <strong>of</strong> sites using graphical type frequency<br />

seriation may or may not produce a correct ordering, to interpret battleship curves as<br />

popularity curves is to misinterpret the seriation schema as reflecting absolute abundance<br />

<strong>of</strong> types over time. (McNutt 1973:51, 58). McNutt considers frequent errors <strong>of</strong> this sort<br />

to be due to casual use <strong>of</strong> the word “frequency”(relative abundance might be a better<br />

term).<br />

Dunnell points out that the extent to which types or attributes form such battleship<br />

curves can be used in dialogue with reclassification to develop a classificatory scheme<br />

(Dunnell 1970:309). Selecting attributes which form battleship curves and produce a neat<br />

seriation is key in Dunell’s view to identifying those aspects <strong>of</strong> variability which are<br />

stylistic.<br />

<strong>The</strong> most widespread seriation technique in the 70s and 80s was to construct an<br />

assemblage similarity matrix using attribute/type counts (converted to relative frequencies<br />

46


summing to 100% for all units), and the Robinson coefficient <strong>of</strong> similarity. It is important<br />

to note that there is no control or weighting for assemblage sample size built into the<br />

resulting similarity matrix, and that, as for graphical frequency seriation, the resulting<br />

pattern <strong>of</strong> site similarity may contain spurious observations from mixed contexts or small<br />

samples. <strong>The</strong>se may be less obvious than in graphical frequency seriation (where the<br />

<strong>of</strong>fending units would be unlikely to fit well anywhere in the series). Fortunately methods<br />

exist for evaluating the goodness <strong>of</strong> fit <strong>of</strong> a matrix sort using this sort <strong>of</strong> data, and seriation<br />

using Robinson coefficient <strong>of</strong> similarity, followed by matrix sorting or cluster analysis has<br />

been successful in many instances.<br />

Lack <strong>of</strong> fit in the sort order can be as a result <strong>of</strong> inadequate sample size, or<br />

representation <strong>of</strong> an anomalous occupation span, possibly <strong>of</strong> material from temporally<br />

separate occupations, but not necessarily so (de Barros 1982, Dunnell 1970: 312). <strong>The</strong><br />

trouble in Oceania is that usually, due to poor sampling, one must put up with a certain<br />

jaggedness in the seriation or wind up with no seriation at all, the corpus <strong>of</strong> sites for the<br />

Lapita period at least being too small to allow fussiness in most cases, when dealing with<br />

Lapita sites from a small area.<br />

McNutt considered that both graphical type seriation and similarity matrix<br />

approaches such as Robinson matrices should be regarded as techniques that work<br />

sometimes rather than as having any methodological validity, and bemoaned the lack <strong>of</strong><br />

caution in their application by the 1970s in contrast to a commended caution displayed by<br />

earlier practitioners (McNutt 1973:58-60). <strong>The</strong>se comments, although not directed at<br />

Oceania similarity matrix-based seriations, seem pertinent here, as will be argued with<br />

detailed reference to specific examples in Chapter 2.<br />

<strong>The</strong> predominant seriation technique in archaeology today, especially in Europe,<br />

is Correspondence Analysis (CA), a multivariate data-exploratory technique suitable for<br />

presence absence data or counts. When used for counts, some control for sample size is<br />

present, as both attribute/type total sample size and the total count for each unit are used<br />

47


to weight data. Attribute/type covariation is controlled for by calculating summary<br />

variables, which are easily plotted in two or three dimensions. Diagnostic information<br />

showing the contribution <strong>of</strong> various attributes/types and units to the summary variables is<br />

available (unlike cluster analyses <strong>of</strong> similarity matrices). As for graphical frequency<br />

seriation and Robinson seriation, if non-chronological variation is input (e.g. spatial and<br />

functional variation) the summary variables obtained may be strongly influenced by these.<br />

Control for spatial variation is possible as input files can include geographic coordinates,<br />

and diagnostic statistics make it clear to what extent variation across space is structuring<br />

the seriation.<br />

Multivariate ordination approaches such as principle components analysis (PCA),<br />

correspondence analysis(CA), principal coordinates analysis and multidimensional<br />

scaling(MDS) have percolated into Oceanian ceramic analysis over the last decade. PCA<br />

and CA are generally more easily evaluated than cluster analyses (Shennan 1997:253, 297-<br />

298). Ordination techniques in general give an indication <strong>of</strong> the relationships between<br />

ceramic variables and units <strong>of</strong> analysis (e.g. site assemblages). <strong>The</strong>y can suggest whether<br />

there are any summary trends in the data. PCA is generally more suited to numeric data,<br />

while CA is better for counted data, although PCA is sometimes used for this purpose.<br />

Seriation is nowadays almost always done using CA (Shennan, 1997 342), on counts <strong>of</strong><br />

types/attributes in the case <strong>of</strong> ceramic refuse, and on presence-absence <strong>of</strong> features in the<br />

case <strong>of</strong> grave-goods or architecture. CA as available in the Bonn Archaeological S<strong>of</strong>tware<br />

Package (BASP or WinBASP on the internet) can output comprehensive diagnostic<br />

statistics to identify the relative contributions <strong>of</strong> the various variables (including geographic<br />

coordinates <strong>of</strong> the samples) and units to the results.<br />

Principal coordinates analysis and MDS attempt to find the structure in a set <strong>of</strong><br />

proximity measures between objects. Results from these analyses are not as easy to<br />

evaluate as those from PCA or CA, as they operate on a similarity matrix rather than on<br />

raw data or data that has been transformed or standardized in some way (Shennan<br />

48


1997:348-349). A key part <strong>of</strong> the MDS method is that it has the advantage <strong>of</strong> providing<br />

an objective indication <strong>of</strong> the number <strong>of</strong> major summary trends that are significant, through<br />

the measure stress. One advantage <strong>of</strong> retaining the use <strong>of</strong> a similarity matrix is that a mix<br />

<strong>of</strong> metric and nominal variables can be used to construct the matrix, particularly useful if<br />

one has metric data on vessel form or the scale <strong>of</strong> decoration.<br />

Occurrence Seriation and Potsherd Samples:<br />

Occurrence seriation in its graphical form arranges samples so that the occurrence <strong>of</strong><br />

criteria is continuous or as near so as possible. <strong>The</strong> usual interpretation is that this<br />

represents an ordering <strong>of</strong> assemblages in time, assuming that any particular attribute was<br />

continuously discarded through time. To be valid this requires, like frequency seriation,<br />

that the attributes used in the seriation do not have an intermittent occurrence, and assumes<br />

control for contemporaneous variation across space and in relation to vessel/site function.<br />

Green’s seriation <strong>of</strong> Lapita sites (Green 1978) and Graves’ and Cachola-Abad’s<br />

seriation <strong>of</strong> architectural features in Hawaii (Graves & Cachola-Abad 1996) are <strong>of</strong> this<br />

type (O'Brien & Lyman 2000:120-121), but there are fundamental differences between the<br />

data used in these two examples which have serious consequences for ceramic seriation.<br />

A tradition <strong>of</strong> occurrence-based assemblage grouping approaches to ceramic seriation has<br />

been a persistent thread in Oceanic pottery analysis, usually using the Jaccard coefficient<br />

or simple-matching coefficient to create similarity matrices (Anson 1983, Best 1984,<br />

Egl<strong>of</strong>f 1971, Frost 1974, Green 1978, Irwin 1972, 1985, Summerhayes 2000a, Wickler<br />

1995, 2001). In the wider archaeological world occurrence seriation has been largely<br />

avoided by ceramicists since Dempsey and Baumh<strong>of</strong>fs initial proposal (O'Brien & Lyman<br />

2000:119), except as a technique for ordering grave-lots. <strong>The</strong> reasons for this avoidance<br />

are firstly that “chronologies based on the presence-absence <strong>of</strong> types or attributes will<br />

never be as accurate as those based on their relative frequencies (Le Blanc 1975).” Le<br />

49


Blanc noted the applicability <strong>of</strong> presence-absence seriation to grave-lots, where<br />

assemblages represent small intentional sets <strong>of</strong> attributes rather than time-accumulated<br />

discard from settlements. Secondly, and perhaps more importantly, the notion that<br />

occurrence seriation is less sensitive to variation in sample size than frequency seriation is<br />

questionable. Sparse data (a lot <strong>of</strong> types/attributes, each with few members) is where<br />

occurrence seriation shines, hence use with grave lots and architectural features. This<br />

assumes though that the sparse data is a complete representation <strong>of</strong> the unit to be<br />

seriated. This point will be more fully discussed below in relation to some archaeological<br />

seriations in Oceania.<br />

In Oceania, for the Lapita period, we have yet to discover a landscape dotted with<br />

generous numbers <strong>of</strong> Pompeii-like yet easily detectable and accessible ceramic vessel<br />

samples, and must be satisfied with a less stringent set <strong>of</strong> limits within which seriation can<br />

be considered to work. Regional site samples can be expected to be on the small side <strong>of</strong><br />

adequate, with assemblage vessel numbers lower than might be ideal, and durations <strong>of</strong><br />

occupation largely guessed at rather than tightly defined. Compounding these preservation<br />

and mixing problems is the problem <strong>of</strong> sparse data (discussed above), where in some<br />

classificatory schemes, a large number <strong>of</strong> types/attributes exist, each with few occurrences,<br />

in only a subset <strong>of</strong> sites. This sparseness may arise through splitting approaches to<br />

classification/grouping, in combination with the nature <strong>of</strong> ceramic production (lack <strong>of</strong><br />

standardization), and, also, as a result <strong>of</strong> the sampling processes acting to make the<br />

archaeologist’s sample a fraction <strong>of</strong> the pottery discarded at the location in the past in<br />

many cases.<br />

In response to the problem <strong>of</strong> sparse data, an approach commonly taken in<br />

Oceania has been to apply a form <strong>of</strong> occurrence seriation, by recording the<br />

presence/absence <strong>of</strong> attributes in units/sites, and by creating a matrix <strong>of</strong> similarities between<br />

units with a similarity coefficient suited to this type <strong>of</strong> data, and then performing either a<br />

computerized sort or a cluster analysis on the similarity matrix. <strong>The</strong> two similarity<br />

50


coefficients commonly used on this sort <strong>of</strong> data are the simple matching coefficient and the<br />

Jaccard coefficient. Of these, the simple matching coefficient regards negative matches as<br />

significant, and is thus less suited to sparse data, while the Jaccard coefficient disregards<br />

negative matches entirely, and is therefore better for this type <strong>of</strong> data (Shennan 1997:228).<br />

<strong>The</strong> widespread application <strong>of</strong> these techniques in Oceanian archaeology pays<br />

insufficient attention to the sample-richness problem for this type <strong>of</strong> sparse occurrence data<br />

from sherd samples (but see Anson 1987, Green 1978, Kirch 1987a). Whether or not a rare<br />

attribute occurs in an assemblage <strong>of</strong> sherds may well be a sampling effect, even for<br />

presence-absence data. Where sites/units vary greatly in sample size, spurious dissimilarity<br />

may result through size-related differences in sample richness.<br />

Consider a hypothetical case (Table 1), <strong>of</strong> two assemblages are drawn from<br />

identical populations, differing only in sample size. All types are represented in the large<br />

sample, but many are missing from the small. Calculation <strong>of</strong> a Jaccard coefficient <strong>of</strong><br />

similarity would result in the following data (where a=positive matches, b= present in the<br />

small assemblage but not the large, c=present in the large assemblage but not the small):<br />

a=3; b=7; c=0. Jaccard coefficient is calculated as a÷(a+b+c) or 3÷(3+7+0)=0.3, when<br />

ideally these being samples from the same thing should show high similarity (s<br />

approaching 1.0). In this example sample size is structuring the similarity coefficient to a<br />

greater degree than behavioural similarity. It can be seen that although the Jaccard<br />

coefficient is suited to sparse data, it assumes a complete representation <strong>of</strong> behaviour in<br />

the sparse data, and is not suited to assemblages which comprise incomplete samples <strong>of</strong><br />

a behavioural unit. Using a technique adapted to sparse data does not correct for<br />

inadequate sample <strong>of</strong> sites or inadequate (or unsaturated cf. Kintigh 1984) within-site /<br />

unit assemblage size (see also Anson 1983:54, Kirch 1987a, Specht 1977). This example<br />

illustrates a major potential problem with the application <strong>of</strong> the Jaccard coefficient to<br />

ocurrence seriation <strong>of</strong> Oceanic ceramic data, which would not arise with data from grave<br />

51


lots or architectural features.<br />

Phyletic Seriation:<br />

Phyletic seriation assumes heritable continuity, without specifying a rule <strong>of</strong> developmental<br />

direction, by emphasising evolutionary change <strong>of</strong> a chosen attribute over time. Seriation<br />

<strong>of</strong> Lapita face designs for example (Spriggs 1990) involves an assumption <strong>of</strong> the evolution<br />

<strong>of</strong> a single design over time, although Spriggs tended more towards evolutionary seriation<br />

in that instance as the direction <strong>of</strong> change was assumed to be from initial representation<br />

<strong>of</strong> faces to a later stylized geometric remnant. A more recent example <strong>of</strong> phyletic seriation<br />

<strong>of</strong> Lapita designs is more sophisticated, in that it constructs multi-lineal evolutionary series<br />

for three different aspects <strong>of</strong> Lapita design (Ishimura 2002). Phyletic seriation is about how<br />

change occurs, how one attribute becomes another, where frequency or occurrence<br />

seriation use arbitrarily defined attribute classes to freeze-frame change for measurement<br />

purposes. Mead also discussed phyletics <strong>of</strong> motif relationships by speculating that some<br />

motifs were developments <strong>of</strong> or elaborations <strong>of</strong> others (Mead 1975:37)<br />

Table 1: Sampling, assemblage richness, and the Jaccard coefficient (p=present, a=absent).<br />

Attribute 1 2 3 4 5 6 7 8 9 10<br />

Small<br />

sample<br />

Large<br />

Parent<br />

Popn.<br />

p a a p a a p a a a<br />

p p p p p p p p p p<br />

52


Seriation in Oceania:<br />

<strong>The</strong> best examples <strong>of</strong> local regional sequence building in Oceania have at their cores a large<br />

sample <strong>of</strong> collection sites, representing a good percentage <strong>of</strong> the total ceramic diversity<br />

within a small survey region. Additionally, they all use graphical attribute seriations and<br />

stratigraphic superposition to check the computer matrix results. Most, but not all, use<br />

ordered similarity matrices for their seriations, based on attribute frequencies and a<br />

Robinson coefficient <strong>of</strong> similarity.<br />

<strong>The</strong> Frost/Irwin numerical taxonomy approach was typified by a similarity grouping<br />

analytical mode. In analysing his Shortland data, for example, Irwin adopted Clarke’s<br />

proposition (Clarke 1968:512-549) that there is a functional convergence between<br />

occupation sites that imposes some control on functional variation, and that relative<br />

similarity between archaeological assemblages therefore indicates closeness in time (Irwin<br />

1972:83-85) (This assumes invariant site function and complete representation <strong>of</strong> the<br />

properties <strong>of</strong> a breakage population). Thus if a similarity matrix is ordered so that an ideal<br />

pattern emerges, the sites are arranged in order <strong>of</strong> chronological distance. Irwin noted the<br />

possibility <strong>of</strong> contemporaneous spatial variability creating such a pattern in the similarity<br />

matrix, but made no specific mention <strong>of</strong> functional variability, which, as vessel form data<br />

were included, could be seen as problematic. Irwin felt that because site surface collections<br />

were quite large and <strong>of</strong> “not especially long duration” (he had culled them <strong>of</strong> material<br />

thought to be intrusive from different periods prior to analysis) there was little likelihood<br />

<strong>of</strong> anything other than temporal variation emerging in his patterning.<br />

Irwin, citing Frost’s Fiji work, used the simple matching, Jaccard and Robinson<br />

coefficients <strong>of</strong> similarity to perform sorted similarity matrix seriations. <strong>The</strong> first two<br />

coefficients were used with occurrence data, while the Robinson coefficient is a<br />

frequency-based measure <strong>of</strong> similarity. Green (citing Irwin) suggested that occurrence<br />

seriation using the Jaccard coefficient was beneficial in reducing sample size differences<br />

(Green 1978), but this is incorrect: the technique is suitable for the analysis <strong>of</strong> sparse data<br />

53


(discussed in more detail below),which is not the same thing. <strong>The</strong> use <strong>of</strong> these two<br />

different types <strong>of</strong> seriation on the same samples has been widely followed in Pacific<br />

archaeology (Anson 1983, Best 1984, Green 1978, Wickler 2001). <strong>The</strong>se techniques<br />

commonly, and not unexpectedly, produce conflicting ordering <strong>of</strong> sites in the resulting<br />

similarity matrices, requiring a decision as to the “correct” ordering, and encouraging a<br />

variety <strong>of</strong> matrix sorting or clustering techniques, all <strong>of</strong> which can be decidedly divorced<br />

from the original archaeological data, and difficult to evaluate.<br />

Irwin’s study <strong>of</strong> a large number <strong>of</strong> late-prehistoric sites within the local Mailu<br />

region analyzed 28 contextual assemblages, some from excavation zones, some from<br />

surface collections, and constructed a sequence using sorted similarity matrices, graphical<br />

attribute frequency seriation and superposition chronological inference (Irwin 1985:118-<br />

162). Similarly, Egl<strong>of</strong>f made use <strong>of</strong> attribute-frequency sorted similarity matrices <strong>of</strong> a large<br />

number <strong>of</strong> surface sites and some mound excavations in the Collingwood Bay region, and<br />

also made use <strong>of</strong> graphical attribute seriations (Egl<strong>of</strong>f 1979:47, Figs 17-19). Egl<strong>of</strong>f’s<br />

sequence ran from the undated Lapita-like “Group P” ceramics to modern. Best applied<br />

Irwin’s method to construct a ceramic sequence for Lakeba in Fiji running from Lapita to<br />

modern. Both Best and Irwin controlled for sample size in their seriations by omitting<br />

smaller assemblages (under 100 sherds in Irwin’s case), but neither discuss quantification<br />

<strong>of</strong> sample size in any depth, or methods by which mixed or long-occupation assemblages<br />

can be identified, an area in which both Specht and Bedford have pointed out the value <strong>of</strong><br />

stylistically homogenous sampling units (Bedford 2000, Specht 1969). In Bedford’s case<br />

it is arguable that some <strong>of</strong> the lack <strong>of</strong> diversity is related to small sample sizes in the lower<br />

levels <strong>of</strong> rockshelter test pits on Malekula, while Specht was referring to larger samples<br />

surface-collected from abandoned village sites on Buka.<br />

Best, like Irwin, used attribute frequencies to calculate a Robinson matrix <strong>of</strong><br />

similarity, and cluster analysis to create a seriation <strong>of</strong> excavation and surface collection<br />

units (Best 2002:18-20). A particular feature worthy <strong>of</strong> note is the inclusion in his<br />

54


seriation diagram <strong>of</strong> a “dead” column tracking the percent decorated, which shows six<br />

peaks <strong>of</strong> decoration through his series, rather than a unimodal history, illustrating the<br />

dangers <strong>of</strong> using a recurring or fluctuating attribute for seriation. Best shows sixty units<br />

in his Robinson matrix, and thirty in his graphical seriation, with twenty-three attributes<br />

shown, illustrating the desirability <strong>of</strong> having a reasonably large matrix <strong>of</strong> types/attributes<br />

and units for seriation, to lessen the possibility <strong>of</strong> random repeats <strong>of</strong> stylistic attribute<br />

combinations over time. It is important to note that in Best’s case and in Irwin’s, that<br />

stratified sequences were available, which aided the identification <strong>of</strong> temporally sensitive<br />

attributes used in seriations.<br />

Selection <strong>of</strong> Variables in the Absence <strong>of</strong> Stratified Sequences:<br />

How best to structure classificatory systems to meet chronological aims is seldom<br />

discussed in archaeology generally. Some attributes/variables may fluctuate over time and<br />

space (the relative percentages <strong>of</strong> decorated vessels versus undecorated vessels for<br />

example), or may vary in relation to function (vessel form for example) which may in turn<br />

affect decoration (Shepard 1963:260-261). <strong>The</strong> danger in analyses that utilize a large<br />

number <strong>of</strong> variables/attributes in an a-theoretical manner is that the signal they are after<br />

(temporal variation for example) will be obscured by other types <strong>of</strong> variation, such as<br />

functional, spatial or sample-size variation. <strong>The</strong>se problems are lessened in CA and PCA,<br />

where it is relatively easy to discover if spatial or sample-size variables are contributing to<br />

the results, but the golden rule in seriation analysis according to Dunnell is careful selection<br />

<strong>of</strong> types or attributes that vary with time (Dunnell 1970). Some sorts <strong>of</strong> attributes carry<br />

higher risk than others, and can usefully be omitted from analysis when the aim is<br />

chronological. It may be possible as a result <strong>of</strong> such an omission to show, eventually, that<br />

such variables do vary over time, but the converse approach, including all variables, may<br />

result in a mishmash <strong>of</strong> variability that might be difficult to disentangle, particularly for<br />

similarity matrix / clustering approaches in which it may be difficult to ascertain which<br />

55


variables are principally responsible for the series obtained.<br />

Just what is it that one’s units <strong>of</strong> classification are measuring? This is one among<br />

a number <strong>of</strong> reasons why validation <strong>of</strong> cluster analyses <strong>of</strong> similarity scores is problematic.<br />

<strong>The</strong> Mead/Donovan and Siorat approaches have such theory, and seek ideas <strong>of</strong> the<br />

decorators regarding the structure (and sequence in the case <strong>of</strong> Siorat) <strong>of</strong> decoration<br />

(Mead 1975:37, Siorat 1990). In regard to the Mead/Donovan and Siorat approaches, if<br />

past ideas <strong>of</strong> structure is their objective, one must question whether there is likely to be a<br />

strong shift in such structures over time alone, rather than over space or in relation to<br />

functional variability. Clark also provides some discussion <strong>of</strong> the theoretical basis <strong>of</strong> his<br />

attribute selection (Clark 1999:65), and Wickler and Anson are both forthright on the<br />

difficulties <strong>of</strong> separating functional and temporal variability. Summerhayes advances a<br />

specific theory <strong>of</strong> ceramic temporal-functional variability (Summerhayes 2000c),<br />

concluding that function varied over time in the West New Britain data.<br />

<strong>The</strong> review <strong>of</strong> seriation method and theory above sets the stage for a critical review<br />

<strong>of</strong> constructions <strong>of</strong> Lapita temporal variability in Chapter 2, and for the analysis <strong>of</strong><br />

variability and seriation <strong>of</strong> Roviana data in this thesis. Frequency seriation, using attribute<br />

abundances and correspondence analysis, has the potential to yield a high-resolution<br />

chronology if samples are fit for the task and time-sensitive attributes are chosen.<br />

Correspondence analysis negates the argument that occurrence seriation is better for<br />

unsaturated samples, since correspondence analysis is weighted by site/attribute sample<br />

size. <strong>The</strong> quantification issue raises its head here though, as sample size must be assessed<br />

using a measure which is suited to the completeness characteristics <strong>of</strong> the sample.<br />

Occurrence seriation is unlikely to match the chronological resolution <strong>of</strong> frequency<br />

seriation for potsherd samples, being more suited to sparse data such as grave goods or<br />

architectural features, where the number <strong>of</strong> attributes <strong>of</strong> a sampling unit are large and the<br />

membership <strong>of</strong> each type/attribute is small, and where the sparse data is a complete<br />

representation <strong>of</strong> each unit such as a set <strong>of</strong> excavated grave goods, rather than a sample.<br />

56


I have argued above for the importance <strong>of</strong> sample evaluation as a first step in<br />

seriation: sample sizes for sites and units are one side <strong>of</strong> this issue (including the difficult<br />

problem <strong>of</strong> the number <strong>of</strong> independent observations that comprises the sample), but<br />

another difficult area is the assumption <strong>of</strong> heritable and historical continuity in the overall<br />

sample. It is important to try to identify sampling gaps or lack <strong>of</strong> heritable continuity<br />

between units or types. Phyletic analysis has a role to play here. Phyletic seriation is a<br />

mode <strong>of</strong> analysis that can usefully be combined with frequency seriation. When a series is<br />

constructed, phyletic analysis allows the investigation <strong>of</strong> the nature <strong>of</strong> the transformations<br />

that occur to make the series. How does one attribute transform into another? Phyletic<br />

analysis in turn leads on to “why” questions: if the nature <strong>of</strong> change can be understood, the<br />

explanation <strong>of</strong> change can be investigated. In Chapter 12 such a combination <strong>of</strong> frequency<br />

seriation (using CA) and phyletic seriation is attempted.<br />

This review has noted the importance <strong>of</strong> identifying the sorts <strong>of</strong> attributes in the<br />

data (whether functional or stylistic, within a Dunnellian theoretical framework- explained<br />

in the following section). <strong>The</strong> link between site/artefact properties and function is by no<br />

means simple, and further review <strong>of</strong> this topic is warranted.<br />

Form and Function: a Review :<br />

Introduction:<br />

Dunnell turned to evolutionary theory to explain why seriation worked, suggesting that<br />

style and function were fundamentally dichotomous (Dunnell 1978). In this view style is<br />

by definition that which is subject to evolutionary drift, while function is that which is<br />

controlled by selection (Dunnell 2001). Functional variation is thus a poor tool for<br />

seriation, as function may be stable or may fluctuate over time with the appearance <strong>of</strong><br />

analogous similarities (adaptations or evolutionary convergences), where style, by<br />

contrast, varies stochastically over time by definition. In Dunnell’s definition <strong>of</strong> style and<br />

57


function, form can have functional and/or stylistic components, as can decoration.<br />

Separating stylistic variation from functional variation is simple in Dunnell’s view: stylistic<br />

variation will seriate well (in battleship curves, in the traditional graphical form <strong>of</strong><br />

frequency seriation), while functional variation will not. Dunnell felt this explained why<br />

proponents <strong>of</strong> seriation in the culture-historical era had been able to create successful<br />

seriations through trial and error, trialling various classifications in a materialist manner<br />

until the desired battleship pattern (a seriation) was obtained, and successfully isolating<br />

chronological variation by this means (in conjunction with testing by other forms <strong>of</strong><br />

dating).<br />

Dunnell’s “method” (my term) assumes that sample sizes from sites, the number<br />

<strong>of</strong> sites in the sample, and the level <strong>of</strong> comparability <strong>of</strong> site occupation spans will be<br />

sufficient to create an even battleship pattern (with exclusion <strong>of</strong> sites that don’t fit due to<br />

small sample sizes or anomalous occupation span). As stated above, this is seldom the case<br />

in Oceania: neither can the Roviana data support this assumption, as will be seen in<br />

Chapter 12. Given a patchy archaeological sample, other methods <strong>of</strong> distinguishing<br />

functional variation from stylistic, prior to seriation, are needed.<br />

Techno-function refers to utilitarian aspects <strong>of</strong> an artefact’s use, as opposed to<br />

soci<strong>of</strong>unction or ide<strong>of</strong>unction (Skibo 1992:33-34). Skibo makes a further division <strong>of</strong><br />

function into intended function and actual function, with intended function being reflected<br />

in vessel design, and actual function reflected in use alteration. <strong>The</strong>ories <strong>of</strong> vessel<br />

form/function correlation (Rice 1987: 211, Smith 1985) thus <strong>of</strong>fer an opportunity to survey<br />

the total Roviana assemblage from the perspective <strong>of</strong> intended techn<strong>of</strong>unction. If intended<br />

vessel techn<strong>of</strong>unction can be attributed to ceramic forms, then function can be controlled<br />

for to some extent by constructing seriations only within functional classes. <strong>The</strong> situation<br />

is complicated in archaeology by the fragmentary state <strong>of</strong> recovered vessels, from which<br />

forms are inferred.<br />

<strong>The</strong> evolutionary concept <strong>of</strong> exaptation (Gould & Vbra 1982) could usefully be<br />

58


applied to the relationship between pottery form and function, allowing vessels <strong>of</strong><br />

multipurpose forms, and allowing uses to which the form is adaptable, but not adapted. It<br />

might be argued that there is no fixed vessel function, that artefact function is diverse and<br />

unpredictable from form and therefore determination <strong>of</strong> vessel function is an impossibility,<br />

but there are a number <strong>of</strong> factors working in the archaeologist’s favour that generate the<br />

expectation <strong>of</strong> a correlation at least between form and function. Firstly, a substantial<br />

ethnoarchaeological literature has appeared in the last twenty years on vessel use life, and<br />

suggests that frequency <strong>of</strong> use and type <strong>of</strong> use correlate strongly with use life (David 1972,<br />

De Boer 1983, Jensen et al. 1999, Longacre 1991, Mills 1989, Shott 1989b, 1996, Tani<br />

& Longacre 1999, Varien & Mills 1997, Varien & Potter 1997). <strong>The</strong> more frequently a<br />

vessel is used the more likely it is to be broken. Cooking and serving are a frequent<br />

activities, and representative settlement ceramic assemblages are likely to be dominated by<br />

utilitarian cooking and serving vessels (Varien & Mills 1997:147). Serving vessels, used<br />

frequently but not subjected to thermal stress are likely to have longer use life, and be less<br />

common in assemblages than would be the case if they were subjected to thermal stress,<br />

while storage and ritual vessels are likely to have even longer use lives, and be less<br />

common in assemblages, even if present in abundance during settlement occupation in the<br />

past.<br />

This formational bias towards cooking and serving vessels in assemblages (Varien<br />

& Mills 1997:148), is good news for seriation, as this bias creates for the archaeologist a<br />

measure <strong>of</strong> control for function, provided the sample obtained is representative <strong>of</strong> the<br />

settlement site as a whole (likely in the case <strong>of</strong> surface scatters exposed over a wide area,<br />

but more difficult to support in the event that one small excavation square is taken as<br />

representative <strong>of</strong> the whole, or if “site” function varies between sites and not all sites are<br />

“occupations”).<br />

This does not negate the need for ascribing function to sherds. One needs to know<br />

what sorts <strong>of</strong> vessel forms one has in the assemblage to be confident that such a bias does<br />

59


in fact exist in any particular sample, and further, it would be useful to be able to exclude<br />

some observations relating to non-cooking/serving vessels from consideration where these<br />

are identifiable.<br />

A use alteration perspective suggests that cooking pots are arguably more easily<br />

distinguished from other vessel classes in archaeological assemblages based on cracking,<br />

sooting and abrasion resulting directly from their different use environments (Skibo 1992).<br />

This is clearly not the case for the Roviana assemblages, where sooting was noted only on<br />

one vessel, (parts <strong>of</strong> which were found buried in mud), while for the majority <strong>of</strong> sherds the<br />

vessel surface was either clean or ablated, probably a consequence <strong>of</strong> the intertidal, surface<br />

context <strong>of</strong> deposition. For this reason vessel form seemed to be the best available<br />

information about vessel function. Nor has a use alteration perspective been applied to<br />

Lapita pottery in any <strong>of</strong> the studies reviewed in this <strong>chapter</strong>. Sooting, at least, in Oceania,<br />

seems hard to find.<br />

In the absence <strong>of</strong> use-alteration information, how might a classification <strong>of</strong><br />

sherd/vessel form according to vessel function be achieved? A number <strong>of</strong> measures <strong>of</strong> form<br />

properties <strong>of</strong> ceramic vessels relating to vessel function have been suggested or<br />

investigated by archaeologists, and a range <strong>of</strong> technical issues have surfaced in relation to<br />

the measurement <strong>of</strong> vessel form, examples <strong>of</strong> which are given below.<br />

<strong>The</strong> Culture History Approach <strong>of</strong> the 1940s and Onward:<br />

In southwestern American archaeology, <strong>of</strong>ten the distinction is merely between jars, for<br />

cooking and storage, and bowls, being serving and eating vessels (Skibo 1992:36), an<br />

unsatisfactory generalization without much theoretical/methodological rigour.<br />

“<strong>The</strong> notion that use strongly influences pot morphology is venerable, but<br />

few archaeologists or ethnographers have examined it in a uniformitarian<br />

theoretical framework (Smith 1985:257).“<br />

Rice reviews several “inferred use” classificatory schemes, including height-to-width<br />

vessel proportions, vessel contour classifications, and geometric/volume classifications<br />

60


(Rice 1987: 215-222). All <strong>of</strong> these schemes place a heavy burden on sherd size, which<br />

needs to be large to reconstruct forms, although Rice’s own approach, a vessel contour<br />

approach, is less restricted by this factor than either the vessel proportion approach or a<br />

geometric/volume classification.<br />

<strong>The</strong> Methodological Uniformitarians:<br />

As archaeologists we would like to be able to specify a deterministic link between form and<br />

function, but this is not possible. Rye for example states,<br />

“If it can be established that specific materials are usually correlated with<br />

specific functions, the presence <strong>of</strong> these materials can be used to infer<br />

function (Rye 1981)”<br />

Rye goes on to detail ceramic properties appropriate to different functions. He seems to<br />

be suggesting that he is providing a manual or method for the identification/determination<br />

<strong>of</strong> vessel function, and yet the information cannot be used in this way in the absence <strong>of</strong> a<br />

clear causal relationship. For example, the proposition that rounded forms improve thermal<br />

shock resistance is not matched by any statement that rounded forms indicate cooking as<br />

an intended function. <strong>The</strong>re are examples in the ethnoarchaeological literature, too, that<br />

suggest flat-bottom vessels are suited to cooking functions also (e.g De Boer 1983: 109,<br />

Rice 1987:224-226).<br />

A number <strong>of</strong> studies have attempted to develop methodological uniformitarian<br />

predictors <strong>of</strong> use based on ethnographic data (Henrickson 1990, Henrickson & Mcdonald<br />

1985, Smith 1983, 1985). <strong>The</strong>se provide cautionary documentation <strong>of</strong> a complex situation,<br />

where some classes <strong>of</strong> vessel use are relatively easily identified for whole vessels (e.g. dry<br />

storage: Smith) but where identification <strong>of</strong> other functions based on morphology is<br />

problematic, even at the level <strong>of</strong> whole vessels, and more so when dealing with broken<br />

sherd assemblages.<br />

Smith’s concern was solely with morphological indicators <strong>of</strong> use, in contrast to the<br />

61


more common technical and use-alteration approaches. Smith used Binford’s distinction<br />

between technomic, sociotechnic and ideotechnic aspects <strong>of</strong> culture (Binford 1962) to<br />

argue that a narrow focus on technomic aspects <strong>of</strong> ceramic function was possible without<br />

being diverted by this wider debate on function and style. Smith proposed no<br />

comprehensive theory <strong>of</strong> ceramic vessel form: it was noted that many factors besides use<br />

determine form. Smith allowed pots could have multiple uses, and specified also a<br />

distinction between morphological variables <strong>of</strong> whole vessels (e.g. vessel volume, height<br />

to width ratio, etc) and sherd variables, the latter being properties describing vessel size<br />

and shape at a single point, using an assumption <strong>of</strong> vessel symmetry.<br />

Smith’s sample <strong>of</strong> 39 ethnographic pots, for which eight ethnographic use classes<br />

were listed, was drawn from six ethnoarchaeological/ethnographic studies <strong>of</strong> the American<br />

southwest, which raises obvious sampling issues; her subsequent dissertation incorporated<br />

an enlarged and geographically/culturally more diverse sample (Smith 1983:116-140). <strong>The</strong><br />

small sample did not allow detailed correlations <strong>of</strong> different factors <strong>of</strong> use with specific<br />

morphological features. Her use classes were cooking; dry storage; wet storage;<br />

processing; serving <strong>of</strong> food; individual consumption; transport <strong>of</strong> liquid; and washing<br />

within the pot. Morphological data were reconstructed from illustrations.<br />

Smith digitized pot pr<strong>of</strong>iles with the aid <strong>of</strong> a shadow device, and encoded these as<br />

polynomial expressions <strong>of</strong> curves using the GLM regression procedure <strong>of</strong> SAS.<br />

Discriminant analysis was used to evaluate which variables were <strong>of</strong> use in defining<br />

function, and which sherd statistics (sherds statistics used experimentally were derived<br />

from the form data <strong>of</strong> the same vessel sample by a computer simulation <strong>of</strong> breakage).<br />

Predictive success was best when broader functional categories was used, which<br />

cut out instances <strong>of</strong> multiple-use (her short list was cooking, dry storage, wet storage or<br />

transport <strong>of</strong> liquids, and other). A clear distinction could be made between dry-storage<br />

vessels (low exposed surface per volume unit and small mouths, with large volumes<br />

characteristic to a lesser extent) and other vessels <strong>of</strong> all types. Consumption vessels<br />

62


(bowls) were well discriminated (large orifice relative to capacity, large orifice in absolute<br />

terms, and small volume). Processing, serving and washing vessels (other) had a<br />

moderately large ratio <strong>of</strong> exposed surface to capacity and moderately large orifices, with<br />

a slight tendency towards small volumes. Cookpots showed more variation than any other<br />

category, suggesting a category ill-defined or too broad, or that crucial morphological<br />

constraints were not monitored by the variables <strong>of</strong> the study. Cookpots overlapped<br />

significantly with liquid storage and liquid transport samples, with the liquid storage and<br />

transport vessels tending to have small ratios <strong>of</strong> orifice to volume, with these vessels not<br />

displaying the expected provisions for closure (Smith 1985: correlate 9 on p265). This last<br />

finding will be shown to be significant to the Roviana case in Chapter 8.<br />

Smith concluded that food storage and individual eating should be activities with<br />

reasonable archaeological visibility (without considering breakage rates). She noted the<br />

need to develop this method in the direction <strong>of</strong> archaeological practice, where some <strong>of</strong> the<br />

orientation variables are difficult or impossible to measure on broken sherds. Her efforts<br />

in this direction are promising, but preliminary, and are a long way from establishing a<br />

uniformitarian theory. Her results do suggest caution is required in the hasty assignment<br />

<strong>of</strong> globular jars and globular jars only to cooking functions and broader, shallower vessels<br />

to serving functions. Along with her finding that liquid storage/transport vessels had larger<br />

orifices than expected, this suggests that the cooking/serving/water jar distinction <strong>of</strong>ten<br />

seen in ethnographic terminology (e.g. Kirch/Green and Ross, discussed below) need not<br />

have a clear morphological correlation (see Rice 1987: figure 7.14 for a similar point in<br />

relation to cooking pot morphology).<br />

Henricksen and McDonald (Henrickson 1990, Henrickson & Mcdonald 1985)<br />

discriminated between function in a general sense (undefined) and primary function (also<br />

undefined). No distinction was drawn between intended function and actual function. <strong>The</strong><br />

approach to the discrimination <strong>of</strong> function was less systematic than Smith's study:<br />

discriminant analysis was not used. As in Smith's study, samples were small.<br />

63


Morphological correlates <strong>of</strong> function were found to vary according to factors not available<br />

to the archaeologist, for example, the mode <strong>of</strong> transport <strong>of</strong> water transport vessels had a<br />

pr<strong>of</strong>ound influence on form.<br />

Some <strong>of</strong> their statements verge on “ just-so stories” masquerading as explanation<br />

for morphology, for example cooking pots usually had: “A somewhat restricted mouth to<br />

prevent rapid evaporation from boiling foods (Henrickson & Mcdonald 1985:631)”. <strong>The</strong><br />

degree <strong>of</strong> mouth restriction seems unlikely to have any bearing on the rate <strong>of</strong> evaporation<br />

<strong>of</strong> the contents, as given a constant energy supply via the vessel body, evaporation can only<br />

be substantially reduced by pressurizing the contents, (as in a pressure-cooker), with orifice<br />

diameter potentially having only a slight effect on the condensation rate <strong>of</strong> vapour leaving<br />

the vessel. Restricted orifice is more likely to reduce spillage and to increase form strength<br />

<strong>of</strong> the vessel.<br />

<strong>The</strong> case studies presented in the article made heavy use <strong>of</strong> use-alteration to assign<br />

function; problematic if a distinction is drawn between intended and actual function.<br />

Measurement <strong>of</strong> Vessel Form:<br />

On the subject <strong>of</strong> vessel form measurement (orifice measurement and other curvature<br />

measurement on ceramic vessels), Plog noted a number <strong>of</strong> studies suggesting that<br />

measurement error was possibly a significant source <strong>of</strong> ceramic variability in some<br />

instances, and might obscure significant variation in others (Plog 1985). Plog suggested<br />

a dial-gauge technique for curvature measurement rather than curve-fitting using curvature<br />

templates, although he noted a number <strong>of</strong> problems with this, most notably the need to<br />

average a number <strong>of</strong> measurements along a curve to establish curve variability. Curvature<br />

measurements in the Roviana study were made using a custom-made set <strong>of</strong> brass curve<br />

templates, rather than the usual diameter chart (Chapter 4), which had some advantages,<br />

not least <strong>of</strong> which was using curve fitting to determine measurement points at pr<strong>of</strong>ile<br />

corner points such as necks, not easily done using Plog’s dial-gauge method.<br />

64


In the review <strong>of</strong> the Lapita ceramic series in Chapter 2, vessel form is an element<br />

<strong>of</strong> some <strong>of</strong> the chronological propositions critiqued there, and uncertainties in the<br />

measurement <strong>of</strong> vessel form are pointed out. Those criticisms are self-explanatory, and<br />

therefore methods <strong>of</strong> vessel form measurement will not be reviewed in detail here.<br />

Similarly, details <strong>of</strong> measurement methods used in analysis <strong>of</strong> the Roviana ceramics are<br />

given in detail in Chapter 4.<br />

Spatial Variation in Form:<br />

<strong>The</strong>re is some evidence from ethnoarchaeology that vessel form can vary spatially across<br />

quite short distances, with no functional difference observable. Ethnoarchaeology <strong>of</strong><br />

Kalinga pottery identified spatial variation in the form <strong>of</strong> pottery over about 10km within<br />

a single functional class (rice-cooking pot) (Longacre 1991:106-108). <strong>The</strong> variation is<br />

similar to the distinction between s<strong>of</strong>t-shouldered globular vessels and hard-shouldered<br />

globular vessels in the Roviana assemblages detailed in Chapters 4 and 8, so it seems that<br />

there can be substantial production variability for the same functional class within a<br />

smallish region. This brings to mind Smith’s caution that vessel form/function categories<br />

be kept broad.<br />

Oceanic Historical Anthropology and Vessel Function:<br />

Green’s substantive uniformitarian theory <strong>of</strong> Polynesian plainware vessel function, based<br />

on ethnographic observation <strong>of</strong> material culture <strong>of</strong> descendant cultures, suggests the<br />

smaller open post-Lapita bowls from Samoa were used for drinking, serving fluid foods,<br />

kava serving, and holding dyes for decorating bark cloth (Green 1974a:129). Medium-<br />

sized bowls could have been used for preparation <strong>of</strong> barkcloth materials, cooking, and food<br />

preparation. Medium and large bowls were equated with Kava receptacles.<br />

Linguistic terms for small cups and boiling have been reconstructed for Proto-<br />

Polynesian (Kirch & Green 2001:136). Similarly, Ross used lexical reconstruction <strong>of</strong><br />

65


Proto-Oceanic terms to infer a Proto Oceanic taxonomy <strong>of</strong> four functional classes <strong>of</strong> vessel<br />

(cooking pots, water jars, bowls and frying “pans”, but noted that identification <strong>of</strong> the<br />

forms <strong>of</strong> these was a task for archaeologists rather than linguists (Ross 1996). <strong>The</strong><br />

simplicity <strong>of</strong> this scheme conflicts in some respects with expectations for middle-level<br />

societies around the world. <strong>The</strong> various storage, processing, ritual and transport functions<br />

observed ethnographically in middle-level societies in various parts <strong>of</strong> the world<br />

(Henrickson & Mcdonald 1985, Smith 1985) suggest Lapita pottery would have had a<br />

more complex range <strong>of</strong> functional classes <strong>of</strong> vessel.<br />

Ross’s tentative assignment <strong>of</strong> Western Lapita form to function has biconical<br />

rounded-base vessels as cooking vessels, with water jars having a narrower orifice<br />

altogether (Ross 1996:figure 2). <strong>The</strong>se are similar to Summerhayes type V vessels<br />

(Summerhayes 2000a:34), Sigatoka type E (Mead et al. 1975:15, figure 1.5) and<br />

Niuatoputapu vessel form 7 (Kirch 1988a:162 and figure 101d, 101e). This is at odds with<br />

Smith's uniformitarian findings, where there was substantial overlap between water jars<br />

and cookpots, suggesting multipurpose forms were possible in this category. Perhaps the<br />

solution here is to regard small-orifice flasks as liquid transport/storage vessels, and larger<br />

biconical forms as potentially multipurpose forms, which could easily fulfil either a<br />

cooking or liquid-transport function.<br />

It has been suggested that highly decorated Lapita ware might be a high-status<br />

exchange item (Kirch 1988b:161, 1997:148), and that the amount <strong>of</strong> decoration in<br />

assemblages might reflect contemporaneous inter-community social relations (Kirch<br />

1997:147, 143-144). <strong>The</strong>re is something <strong>of</strong> a conflict between this model <strong>of</strong> decorated<br />

Lapita as exchange items and another proposed by Kirch <strong>of</strong> decorated Lapita as a<br />

depiction <strong>of</strong> an ancestor <strong>of</strong> an Austronesian house (Kirch 1997: 188-191). While both<br />

models are consistent with the notion <strong>of</strong> pots being tattooed for display/presentation, pots<br />

as ancestors and pots as exchange items seem mutually exclusive, as house-ancestor<br />

evokes in my mind manufacture by a locally-resident member <strong>of</strong> the descent-group, rather<br />

66


than an exchange item (but this is a subjective interpretation) and these could be seen as<br />

competing models, or models specific to particular vessel types. An ethnoarchaeology <strong>of</strong><br />

the manufacture and exchange <strong>of</strong> material representations <strong>of</strong> Austronesian house-ancestors<br />

might clarify this point. At present my objection is not founded on any good data.<br />

Neither <strong>of</strong> these models <strong>of</strong> the function <strong>of</strong> Lapita pottery have been discussed in<br />

relation to formation theory regarding vessel function and breakage rates. If decorated<br />

Lapita is high-status exchange ware with an occasional serving function, breakage rates<br />

could be expected to be low, whereas if the decorated bowls and dishes were in frequent<br />

use breakage rates would be higher, and frequency in assemblages would reflect this.<br />

Similarly, if pots symbolising house-ancestors were curated in a quiet part <strong>of</strong> the house,<br />

breakage rates could be expected to be extremely low, unless some specific regular<br />

breakage ritual is invoked, and such vessels could be expected to occur only rarely in<br />

assemblages. Use life and discard rated thus provide a test <strong>of</strong> functional interpretations <strong>of</strong><br />

decorated Lapita. High percentages <strong>of</strong> decorated Lapita pottery in assemblages imply a<br />

high relative breakage rate, or perhaps a historically-particular mechanism <strong>of</strong> deposition,<br />

such as abandonment, followed by subsequent breakage <strong>of</strong> a curated behavioural<br />

assemblage <strong>of</strong> pottery (a life-assemblage) (Brenchley & Harper 1998, De Boer 1983). If<br />

we are to model vessel function for Oceanic pottery, we cannot ignore the link between<br />

vessel function, size or frequency <strong>of</strong> use and breakage-rates.<br />

Summerhayes summarized widespread evidence for the loss <strong>of</strong> complex Lapita<br />

forms with dentate decoration, and relatively unchanging persistence <strong>of</strong> simple<br />

undecorated forms in many areas, seeking explanation for these decoration changes in his<br />

Arawe Islands ceramic sequence (Summerhayes 2000c:293-303). Summerhayes saw the<br />

slow rate <strong>of</strong> change <strong>of</strong> the plainwares as related to an “ongoing domestic/utilitarian role”,<br />

with rapid change over time in the level <strong>of</strong> representation <strong>of</strong> dentate-ware in assemblages,<br />

and the shape, motif type and production <strong>of</strong> dentate pottery. I critique Summerhayes'<br />

chronological conclusions in detail in Chapter 2 as being a too-direct reading <strong>of</strong><br />

67


chronology from superposition, with insufficient attention to cultural or natural formation<br />

processes.<br />

<strong>The</strong> high labour-input into dentate decoration was seen by Summerhayes as<br />

suggestive <strong>of</strong> a high-value or prestige “goods”, with the loss <strong>of</strong> this characteristic over time<br />

“equating” to a lessening <strong>of</strong> the social importance <strong>of</strong> these pots. Summerhayes endorsed<br />

Rathje’s contention that the dentate motifs were Austronesian social/ideological signifiers<br />

(Rathje 2000), extending this by suggesting these functioned in an environment <strong>of</strong><br />

continuing interaction between widespread communities, and that loss <strong>of</strong> these social<br />

signifiers indicates changes in the nature <strong>of</strong> interaction between communities. This<br />

explanation is similar to Kirch’s prestige goods explanation, and unlike Kirch’s “house-<br />

ancestor” model, in that the “house-ancestor” pattern is based on modern ethnographic<br />

observation and persists until the present across most <strong>of</strong> the Austronesian world, making<br />

it a poor candidate for a vanished system. Like Kirch, Summerhayes paid no attention to<br />

the link between function, breakage/discard rates, and assemblage composition.<br />

<strong>The</strong>ories <strong>of</strong> the Functional Correlates <strong>of</strong> Rim Form Variation:<br />

Rim form, is a property <strong>of</strong> vessels relatively easily identified from sherds, and worth<br />

considering in more detail from a functional perspective. Rim form incorporates the<br />

attributes <strong>of</strong> the neck (in the case <strong>of</strong> restricted vessels with everted rims), rim and lip,<br />

including the height <strong>of</strong> rim, degree <strong>of</strong> eversion/inversion, rim and neck pr<strong>of</strong>ile curvature,<br />

changes in thickness across the pr<strong>of</strong>ile and neck orifice size. When tested<br />

ethnoarchaeologically, observable rim form variability may bear only a loose relationship<br />

to use (as suggested by Miller 1985:51-74). Nevertheless, if one wishes to avoid forcing<br />

contemporaneous functional rim form variation into a spurious time series, consideration<br />

<strong>of</strong> the possible meanings <strong>of</strong> form variance and some consequent control for rim form<br />

variance in seriation is desirable.<br />

Everted rims can be regarded as multi-functional components <strong>of</strong> vessels. <strong>The</strong>y may<br />

68


add form strength, reduce spillage (Henrickson & Mcdonald 1985:634), allow a tied cover<br />

for storage or fermentation (Henrickson & Mcdonald 1985:632, Smith 1985:263), aid<br />

pouring (Smith 1985:263) and filling, aid stacking (Miller 1985:63, 70, Skibo 1992:77) and<br />

aid lifting when hot by use <strong>of</strong> a torque or tongs (Skibo 1992:66, Smith 1985:263).<br />

Reduction <strong>of</strong> spillage might be best achieved by having a tall vertical or tall inverted rim,<br />

but filling, pouring, stacking and lifting might encourage a more everted form. Heavily<br />

everted and very short rims are probably not the best design for filling, lifting, or<br />

preventing spillage, but may improve functionality for tying covers, pouring and access for<br />

ladling or stirring, and may improve form strength, extending use-life.<br />

In some methods <strong>of</strong> pot manufacture, rim form may be constrained by the plastic<br />

limits <strong>of</strong> the clay-temper mix. Slab-built rims or coil-built rims make little demand on the<br />

plasticity <strong>of</strong> clay during construction. A tall thick, heavily everted rim can be formed from<br />

an arc-shaped slab or by coil building. In contrast, if the entire vessel including the rim is<br />

formed from a single lump <strong>of</strong> clay by the paddle-and-anvil method, there are potential<br />

difficulties in forming a tall and strongly everted rim, as the taller and more everted the rim<br />

the more the clay must be thinned towards the lip, relative to the neck, potentially resulting<br />

in an overly convergent, and thus weak, rim pr<strong>of</strong>ile (thick at the neck, thin at the rim). If<br />

a heavily everted rim is desired for some reason from a pot built by the one-piece method,<br />

the rim may have to be short rather than tall to prevent excessive thinning towards the lip,<br />

resulting in a rolled rim form in the extreme case. If a tall rim is desired in a one-piece pot<br />

(to prevent spillage for example) it may not be practical to have a heavily everted form, as<br />

the lip might split during manufacture or become overly fragile with reduced thickness.<br />

Form and Function: Conclusions:<br />

<strong>The</strong> detailed identification <strong>of</strong> intended use <strong>of</strong> ceramic vessels from archaeological<br />

contexts, relying solely on vessel morphology, is not currently well supported by any<br />

69


established body <strong>of</strong> theory. <strong>The</strong>re are indications from methodological uniformitarian<br />

studies that some broad categories <strong>of</strong> intended use might be identifiable from sherd<br />

morphology in the future, but there are also strong indications from these studies that form<br />

is determined by other contextual factors in addition to intended use, including pottery-<br />

making tradition. While these rare uniformitarian studies are widely cited as providing a<br />

theory <strong>of</strong> the morphological correlates <strong>of</strong> function (for example Sinopoli 1991:84) few<br />

would claim that these correlates are deterministic laws. Some <strong>of</strong> the morphological<br />

correlates <strong>of</strong> function suggested by earlier studies (Ericson et al. 1972, Linton 1944) have<br />

not been borne out by uniformitarian research other than in the most general sense.<br />

This does not, however, negate the virtue <strong>of</strong> controlling for vessel form variation<br />

in seriations, using broad form-categories. While form may vary systematically over time<br />

as a result <strong>of</strong> change in function, in which case a form-seriation may be justified, this<br />

requires chronology to be established by other, independent means, particularly from<br />

Dunnell’s position, that seriation <strong>of</strong> functional variation is a contradiction in terms. Where<br />

the chronology is not known, and seriation is to be the primary means <strong>of</strong> constructing<br />

chronology, control for form is essential, to reduce the possibility <strong>of</strong> incorrectly seriating<br />

contemporaneous or fluctuating vessel functional variation.<br />

Chapter summary and conclusions:<br />

<strong>The</strong> Roviana intertidal archaeological record, investigated as a means <strong>of</strong> assessing whether<br />

Lapita was continuously distributed across Near Oceania or whether there was a huge gap<br />

in the Near-Oceanic Solomon Islands, requires archaeological theory and method<br />

developed for or adapted to the nature and location in the landscape <strong>of</strong> the material under<br />

study. Understanding temporal variability, always a basic task in the archaeology <strong>of</strong><br />

poorly-understood regions, is complicated by the pottery being in the sea. How can this<br />

record be interpreted in behavioural terms? How can a high-resolution chronology be<br />

constructed? How much <strong>of</strong> a sample is needed? How should sample size be quantified?<br />

70


What are the biases present in the samples? Pottery recovered as surface scatters from the<br />

intertidal and reef flat raises a series <strong>of</strong> questions regarding cultural and natural/post-<br />

deposition formation processes, the investigation <strong>of</strong> which can contribute both to<br />

understanding variability in the samples, and to higher-level questions about sample<br />

preservation and visibility, and the implications for the writing <strong>of</strong> Oceanic prehistory.<br />

What is the state <strong>of</strong> preservation <strong>of</strong> the samples? In what taphonomic<br />

circumstances? A summary <strong>of</strong> the physical geography/geology <strong>of</strong> the research region noted<br />

two unusual properties <strong>of</strong> the Roviana landscape, which relate to the preservation <strong>of</strong><br />

surface sites in the sea:<br />

• the Holocene marine transgression has approximately kept pace with tectonic<br />

uplift, as Plio-Pleistocene uplifted barrier reefs are a few metres above sea level<br />

across most <strong>of</strong> the survey region, which is inferred to mean that sea levels have<br />

been relatively stable over recent millennia, falling at least1.5 metres from a high-<br />

stand circa 4500bp, exposing ceramic deposits that were formerly subtidal to<br />

littoral-zone processes, creating lag deposits <strong>of</strong> ceramics, favouring detection by<br />

archaeologists.<br />

• the chain <strong>of</strong> upraised barrier reefs that encircle New Georgia and form a series <strong>of</strong><br />

lagoons, <strong>of</strong> which Roviana is one, shelters the coastline within the lagoons from all<br />

ocean waves, favouring the preservation <strong>of</strong> littoral-zone ceramics, also a factor in<br />

the detection <strong>of</strong> these sites by archaeologists.<br />

<strong>The</strong> extent <strong>of</strong> these effects is investigated in detail in Chapter 5.<br />

Sample acquisition, through the systematic survey and sample collection <strong>of</strong> a<br />

region, is the basis <strong>of</strong> all further analysis. How should the survey region or regions be<br />

chosen, and what methods should be used to survey it/them? How extensive, and how<br />

intensive, should the survey be? How much information is enough for the questions to be<br />

asked <strong>of</strong> the material to be satisfactorarily answered? In Chapter 2 these questions are<br />

investigated, with examples from surveys in Near Oceania reviewed. <strong>The</strong> methods and<br />

71


esults <strong>of</strong> the Roviana surveys are discussed, and some conclusions are drawn regarding<br />

research design.<br />

A review <strong>of</strong> approaches to ceramic classification identified a need for a materialist<br />

approach, where the units <strong>of</strong> classification used derive from an analysis <strong>of</strong> variability <strong>of</strong> the<br />

samples in hand, in a dialectic with the research questions. Temporally sensitive attributes<br />

are needed if a fine-grained chronology is to be constructed, and the identification <strong>of</strong> these<br />

requires a careful analysis <strong>of</strong> the sorts <strong>of</strong> dimensions <strong>of</strong> variability in the data, whether<br />

functional, stylistic, geographic, sample-size-related, or to do with the brokenness or<br />

completeness <strong>of</strong> the sherd samples from which pottery characteristics are inferred. Units<br />

<strong>of</strong> classification that are not sensitive to differences in vessel brokenness are needed, and<br />

units <strong>of</strong> decorative classification need to be identifiable by location on the vessel, to capture<br />

the structure <strong>of</strong> decoration across the vessel, even in broken sherd assemblages. With these<br />

requirements in mind it was concluded that latitudinal bands <strong>of</strong> decoration should be<br />

central in the classification, as these were identifiable by location even for very broken<br />

assemblages, due to location near to the rugged and well-preserved corner points (lips,<br />

necks, carinations) <strong>of</strong> vessels, which also had the advantage <strong>of</strong> boosting sample size.<br />

Chapters 4 and 9 cover this in detail.<br />

Evaluation <strong>of</strong> the recovered ceramic samples is an important step in the study <strong>of</strong><br />

formation processes. Estimating the brokenness and completeness <strong>of</strong> our samples allows<br />

choice <strong>of</strong> appropriate units <strong>of</strong> quantification for evaluating sample size. Comparison <strong>of</strong> the<br />

sample with an inferred breakage population is also possible in many situations, especially<br />

where samples derive from surface sites, or large area excavations. <strong>The</strong> relationship<br />

between the breakage population and the extant deposit can delineate the potential for<br />

taphonomic bias in recovered samples, and provides a method <strong>of</strong> responding to Gosden’s<br />

call for controlled comparisons. Analysis <strong>of</strong> brokenness and completeness is reported in<br />

Chapter 5.<br />

A comparative study <strong>of</strong> wave exposure <strong>of</strong> the Roviana ceramic sites is needed to<br />

72


e able to assess the relative contributions <strong>of</strong> natural and cultural formation processes. In<br />

Chapter 6 the oceanographic setting <strong>of</strong> the sites is described, and an analysis <strong>of</strong> relative<br />

wave exposure is undertaken. <strong>The</strong>se findings form the backdrop to the study <strong>of</strong> other<br />

formation processes. A range <strong>of</strong> formation models are evaluated in Chapter 7 with<br />

reference to the properties <strong>of</strong> sherd assemblages. Spatial analysis as an aid to the<br />

identification <strong>of</strong> formation processes is deferred to Chapter 11, as it is necessary to<br />

formalize units <strong>of</strong> classification <strong>of</strong> materials prior to undertaking spatial analysis.<br />

An extended discussion <strong>of</strong> seriation theory and the need to control seriations for<br />

vessel form variation was given above. In Chapter 8 the details <strong>of</strong> an analysis <strong>of</strong> form<br />

variability are presented, preparatory to seriation analysis. It was also pointed out above<br />

that units <strong>of</strong> classification need to be adapted to the place and period under study, and also<br />

to the range <strong>of</strong> variation present in the sample. For the Roviana materials this means that<br />

the units <strong>of</strong> classification used need to be designed to capture with sensitivity the<br />

transition from a style <strong>of</strong> pottery that is recognizably Lapita to something that is not<br />

Lapita. An analysis <strong>of</strong> variability is needed to translate the initial units <strong>of</strong> description into<br />

units <strong>of</strong> classification appropriate to these aims. Furthermore, units <strong>of</strong> classification need<br />

to be insensitive to differences in vessel brokenness if they are to be useful for<br />

comparisons. Finally, it is beneficial to the analysis <strong>of</strong> relatively plain pottery, in terms <strong>of</strong><br />

the sensitivity to change, if units <strong>of</strong> classification make use <strong>of</strong> the information to be had<br />

from the structure <strong>of</strong> decoration across the vessel. For all <strong>of</strong> these reasons a detailed<br />

exploration <strong>of</strong> ceramic variability is desirable, reported in Chapter 9.<br />

<strong>The</strong> collection sites included lithic items alongside ceramics. While there has been<br />

little discussion <strong>of</strong> these in this introduction, everything written above regarding treating<br />

ceramic sherds as sedimentary particles applies equally to lithic manuports and artefacts.<br />

Analysis <strong>of</strong> lithics in Chapter 10 is made largely from the point <strong>of</strong> view <strong>of</strong><br />

geoarchaeological sourcing <strong>of</strong> raw materials, although a discussion <strong>of</strong> adze forms is<br />

included also. <strong>The</strong> sourcing analysis, which seeks to identify procurement units, is the<br />

73


asis <strong>of</strong> the lithic component <strong>of</strong> analysis <strong>of</strong> intrasite spatial structure in Chapter 11. Spatial<br />

analysis yields information bearing on the identification <strong>of</strong> cultural and natural formation<br />

processes. In Chapter 11 spatial units <strong>of</strong> classification are re-evaluated prior to seriation<br />

analysis. One collection site (Zangana) is split into two spatial units for seriation as a result<br />

<strong>of</strong> the spatial analysis.<br />

<strong>The</strong> extended discussion <strong>of</strong> seriation given above is translated into practical analysis<br />

<strong>of</strong> Roviana ceramic data in Chapter 12. Carbon 14 data and thermoluminescence data are<br />

integrated with seriation data to produce a provisional chronology in need <strong>of</strong> further<br />

independent corroboration and fleshed-out sampling.<br />

Information emerging from these various analytical <strong>chapter</strong>s is pulled together in<br />

Chapter 13. Here the focus returns to the significance <strong>of</strong> the findings regarding Lapita<br />

detectability in Near Oceania. Despite the difficulties <strong>of</strong> constructing a reliable chronology<br />

with the sample in hand, there is also some discussion comparing the Roviana data with<br />

ceramic data from other areas. <strong>The</strong>re is a distinct possibility that some ceramic variants<br />

mixed in with Lapita in other places are strongly separated spatially from Lapita in the<br />

Roviana lagoon early ceramic sample. If this is so, then the approach taken to constructing<br />

chronology, across space rather than by the bedded strata method, is amplifying temporal<br />

resolution beyond what is obtainable from the test-pitting approach <strong>of</strong> the bedded-strata<br />

school.<br />

Ceramic petrographic sourcing has not been discussed in this introduction, and will<br />

not be discussed in any detail in this thesis. Results <strong>of</strong> this sort <strong>of</strong> analysis were interesting<br />

though, but for reasons <strong>of</strong> brevity discussion <strong>of</strong> the implications <strong>of</strong> work either published<br />

elsewhere or appended (appended see Dickinson 2000a, Felgate & Dickinson 2001) is<br />

confined to a short section in Chapter 13, integrating the analysis <strong>of</strong> ceramic tempers with<br />

petrographic analysis <strong>of</strong> lithics.<br />

Before embarking on these analyses <strong>of</strong> variability <strong>of</strong> the Roviana materials though,<br />

it is necessary to review existing literature pertaining to Lapita and the archaeology <strong>of</strong><br />

74


Near Oceania, in the light <strong>of</strong> the methodological review above, to provide ceramic-<br />

chronological context for the Roviana study. <strong>The</strong> particular foci <strong>of</strong> this review in the<br />

following <strong>chapter</strong> will be the temporal resolution <strong>of</strong> existing constructs <strong>of</strong> Lapita temporal<br />

variability, and the level <strong>of</strong> confidence with which these can be accepted as tested<br />

syntheses.<br />

75


CHAPTER 2:<br />

A REVIEW OF CONSTRUCTIONS OF LAPITA<br />

Introduction:<br />

TEMPORAL VARIABILITY<br />

<strong>The</strong> Lapita Ceramic Series (Golson 1971, Green 1978, 1979, 1990, see also Sand 2000)<br />

while not a term in universal use, embodies a concept <strong>of</strong> Lapita ceramics as a time-<br />

transgressive changing system <strong>of</strong> ceramic production/discard comprising regionally varied<br />

sequences within the overall region in which Lapita has been found. Summerhayes<br />

recently suggests a universal series, where there is geographic variation, but where in<br />

some respects ceramics change in parallel through time across the Lapita distribution,<br />

indicating continuing interaction (Summerhayes 2000a, b, 2001, 2002). In this review I<br />

suggest that the security with which a Lapita ceramic series has been defined has been<br />

over-stated in most cases examined, and that formulations <strong>of</strong> a Lapita ceramic series are<br />

mostly working hypotheses rather than confirmed histories <strong>of</strong> ceramic change. This means<br />

that higher-level inferences regarding social processes, such as colonization rates (e.g.<br />

Anderson 2002) are not founded on secure high-resolution chronologies. Getting the<br />

historical facts straight is vital if explanations for change are to be stable and sensible<br />

(Spriggs 2001).<br />

Golson (1971) noted technological continuities between Lapita and post-Lapita<br />

plainware, and in view <strong>of</strong> this evidence for homologous similarity, and also in regard <strong>of</strong><br />

the increasingly varied corpus <strong>of</strong> radiocarbon dates accumulating from Lapita sites,<br />

extended Gifford’s conception <strong>of</strong> Lapita as a widespread pottery time-horizon (Gifford &<br />

Shutler 1956:93-95) to include the notion <strong>of</strong> temporal variability by use <strong>of</strong> the term<br />

tradition, a situation for which Golson suggested the term series as appropriate (Golson<br />

1971:75) (although Gifford did not use the term horizon, he clearly regarded the Watom,<br />

77


Isle De Pins, Site 13 and Sigatoka materials as roughly <strong>of</strong> equal age). Golson’s<br />

understanding <strong>of</strong> “series” was that: “<strong>The</strong> series is made up <strong>of</strong> a set <strong>of</strong> ceramic styles which<br />

are similar and contiguous in either space or time or both (Golson 1971:75)”. Golson’s<br />

“Lapitoid” ceramic series was a hypothetical rather than an “established” detailed<br />

construct, as “...the individual styles that it includes have yet to be isolated, described and<br />

named (Golson 1971:75).”<br />

Green constructed a series for plainware in Samoa, running from circa 800BC<br />

Lapita to Polynesian plainwares by 500BC, and saw parallels in the sequences <strong>of</strong> Fiji and<br />

Tonga (Green 1974b:253), but objected to Golson’s use <strong>of</strong> the term Lapitoid, in spite <strong>of</strong><br />

the evidence for continuity with Lapita, in view <strong>of</strong> the absence <strong>of</strong> definitive (in Green's<br />

definition <strong>of</strong> Lapita) Lapita design motifs on these plain vessels (Green 1974b: 250-251).<br />

He preferred at that time to think <strong>of</strong> Lapita sensu strictu as a style-horizon, characterized<br />

by a wide range <strong>of</strong> elaborate vessel forms, divisible at most into a series comprising Early<br />

Western and Early and Late Eastern Lapita styles, on the basis <strong>of</strong> differences in vessel<br />

shape and the style and frequency <strong>of</strong> decoration (Green 1974b:251). Green saw the demise<br />

<strong>of</strong> Lapita, and then <strong>of</strong> Polynesian plainware, as a functional change, possibly simply<br />

culinary, but also possibly indicating fundamental social changes consequent on settlement<br />

<strong>of</strong> a previously unoccupied island world (Green 1974b:253). Green defined the Lapita<br />

Ceramic Series as<br />

“...assemblages in which various shouldered pots, jars and bowls, as well<br />

as flat-bottomed dishes and plates, occur in association with widely varying<br />

percentages <strong>of</strong> dentate-stamped, notched and incised decoration. <strong>The</strong><br />

decoration consists <strong>of</strong> a catalogue <strong>of</strong> elements and motifs whose<br />

combinations can be listed and compared. In addition, there is a range <strong>of</strong><br />

infrequently decorated bowls <strong>of</strong> simple shapes and varying sizes, plus<br />

several forms <strong>of</strong> rather more frequently decorated sub-globular pots. Site<br />

assemblages <strong>of</strong> Early Western and Early and Late Eastern styles can be<br />

recognized within the series by differences in vessel shape and by the style<br />

and frequency <strong>of</strong> the decoration.” (Green 1974b:251)<br />

78


Green’s 1978 Lapita Ceramic Series:<br />

Green summarized the existing Mead-system analyses as indicating a division into<br />

Western and Eastern Lapita, which both shared a core <strong>of</strong> widespread “early” motifs. <strong>The</strong><br />

Western division was characterized as having a diversity <strong>of</strong> complex motifs, while the<br />

Eastern Lapita division had a relatively poor inventory <strong>of</strong> comparatively simple motifs.<br />

Green defined these motif sets more specifically in his subsequent analysis <strong>of</strong><br />

Lapita motif occurrence (Green 1979), departing substantially from Golson’s and<br />

Specht’s sequence <strong>of</strong> “styles” approach, taking instead an attribute/computer clustering<br />

approach to group similar assemblages in an ordered similarity matrix, very much in the<br />

Irwin mold (Green 1990:37), using Mead-system classificatory units. Although his<br />

primary objective in highlighting differences between Eastern and Western Lapita in this<br />

way was to test his hypothesis that the Fiji/Vanuatu water gap caused an early break in<br />

a unified Lapita interaction system, his paper included a number <strong>of</strong> chronological<br />

statements concerning agreement between his C14 chronology and decorative similarity,<br />

which effectively constitute an occurrence seriation.<br />

At the time Green conceived <strong>of</strong> the sites in his analysis (the Reef-Santa Cruz sites<br />

at least) as slices in time over up to 1000 years <strong>of</strong> Lapita production/discard, and had<br />

pooled C14 ages by site to reflect this view (Green 1976a). Occupation span was not the<br />

focus <strong>of</strong> any extended discussion. <strong>The</strong>se sites are some <strong>of</strong> the best-sampled and dated<br />

Lapita sites on record, and therefore provide an illustration <strong>of</strong> the need for a general shift<br />

in emphasis from the age <strong>of</strong> the “site” to the age <strong>of</strong> the artefact in seeking to construct<br />

high-resolution chronology. This can be done by focusing on the Reef-Santa-Cruz case<br />

where the sampling difficulties <strong>of</strong> a reconnaissance test-pitting approach have been<br />

avoided by systematic surface collection and large area excavations.<br />

Green regarded the closeness <strong>of</strong> sites SZ-8 and Vatcha in the matrix (Green<br />

1978:figure 7) as confirmation <strong>of</strong> the early date <strong>of</strong> both sites, as suggested previously by<br />

the particular archaeologists concerned with those sites (Green/Donovan for SZ-8 and<br />

Frimigacci for Vatcha). He also saw this analysis as reconfirming the close relationship<br />

between RL-2 (RF2) and Watom.<br />

Green presented one further matrix, an analysis <strong>of</strong> the three Reef-Santa Cruz<br />

assemblages (Green 1978:figure 8). He re-stated his 1976 position, that the relative ages<br />

79


Table 2: Reef/ Santa Cruz motif counts as given in Anson 1983.<br />

Site Total Motif count Motif Richness<br />

RF2 841 examples 178 Motifs<br />

RF6 252 examples 79 Motifs<br />

SZ8 627 examples 133 motifs<br />

<strong>of</strong> the sites were known, and saw a high degree <strong>of</strong> motif sharing between these sites as<br />

indicative <strong>of</strong> a strong continuity over an extended period among those motifs restricted<br />

to the Reef-Santa-Cruz area. He saw a different temporal trend in the Reef-Santa Cruz<br />

sites to that noted by Birks and Shaw for the Eastern Lapita sequences, suggesting<br />

instead a trend <strong>of</strong> initial motif efflorescence followed by impoverishment through time,<br />

as evidenced by the motif-poor RF6 sample. Green had applied a correction for sample-<br />

size-related richness using Donovan’s unpublished motif frequencies to assess whether<br />

a motif was absent as a result <strong>of</strong> sampling error or absent-not-present. This assessment<br />

is challenged below, on the basis that allowance for the effect <strong>of</strong> sample-size differences<br />

was probably insufficient.<br />

Despite subsequent revision <strong>of</strong> the C14 chronology (Kirch & Hunt 1988) which<br />

had the three sites <strong>of</strong> indistinguishable age, Green considered that the sequence <strong>of</strong> sites<br />

as outlined by in his (1978) seriation could still be supported on the basis <strong>of</strong> Donovan’s<br />

chronological inferences from the ceramics (Green 1991c). Best has recently taken issue<br />

with this view, suggesting that the diversity <strong>of</strong> motifs in sites is sample-area related (Best<br />

2002:91). Best argued that the chronology <strong>of</strong> the sites should be reversed. Here I present<br />

an alternative view, similar to Best’s in that the “temporal” similarity relations are<br />

suggested to have more to do with sample size than initially thought. I look to Donovan’s<br />

and Parker’s theses for additional information on sample sizes, assemblage brokenness,<br />

and relative frequencies <strong>of</strong> decorative techniques (Table 2, Table 3, Table 4).<br />

Donovan’s hypothesis was that Reef-Santa Cruz Lapita had a decorative<br />

80


Table 3: Sherd counts and MNI assembled from various tables in Parker (1981).<br />

Site Dentate<br />

rims<br />

“richness” that suggested a greater developmental time-depth than had previously been<br />

considered (Donovan 1973:IV). This was consistent with the then C14 chronology, which<br />

had the sites as slices in time over a period <strong>of</strong> many centuries. Assemblages from three<br />

sites were analyzed, RL2 (RF2), thought at that time to date to approximately 900BC,<br />

RL6 (RF6), approximately 600BC and SZ8, undated at that time.<br />

All rims were removed from these assemblages prior to analysis for special study,<br />

and are not included in Donovan’s counts (Donovan 1973:5). <strong>The</strong> rims were later<br />

analyzed by Parker, providing a useful independent quantification <strong>of</strong> relative sample sizes<br />

(Table 3) (Parker 1981). Counts (sherd counts?) <strong>of</strong> decorative techniques were given by<br />

layer by Donovan, but these data cannot be read as sample sizes in the absence <strong>of</strong> any<br />

information on vessel fragmentation, completeness, part representation, or sherd sizes.<br />

<strong>The</strong> relative frequencies <strong>of</strong> decorative techniques were calculated with counts <strong>of</strong> plain<br />

sherds included, which tends to muddle the picture, as plain vs decorated may easily<br />

fluctuate over time rather than follow a steady trajectory, and is sensitive to differences<br />

in brokenness. <strong>The</strong> percentage <strong>of</strong> dentate to incised sherds by site/layer was recalculated<br />

with counts <strong>of</strong> plain sherds excluded (Table 4).<br />

<strong>The</strong>se data do not support any change in relative frequency <strong>of</strong> dentate-stamping<br />

to incised decoration by level in any <strong>of</strong> the three sites (which is not surprising as the sites<br />

had been subject to post-depositional gardening), but more significantly, differences<br />

between<br />

Incised<br />

Rims<br />

81<br />

Impressed<br />

rims<br />

Plain<br />

rims<br />

Total Rim<br />

Sherd<br />

count/<br />

MNI<br />

RF6 43 1 15 2 61/55<br />

SZ8 111 8 10 14 143/?<br />

RF2 848 50 425 100 1423/666


Table 4: Relative proportions <strong>of</strong> dentate and incised, recalculated from data presented by<br />

Donovan (1973).<br />

Dentate<br />

Count<br />

sites are slight. (As presented by Donovan, there were substantial differences in the<br />

relative frequency <strong>of</strong> techniques by site and by layer, but as the recalculation shows, these<br />

have more to do with differences in the percentage <strong>of</strong> plain sherds, which can mean<br />

differences in brokenness between contexts, than with differences in potting behaviour<br />

in the past.)<br />

Donovan thought that marked decorative similarities between the sites were due<br />

to common origin within the same tradition, and strong regionalism (Donovan 1973: 36,<br />

43). <strong>The</strong> possibility that the sites were similar because they were all <strong>of</strong> similar or<br />

overlapping age and that differences arose for reasons other than chronology, was not<br />

considered, because the chronology was thought to be largely known from the C14<br />

evidence.<br />

Dentate % Incised<br />

count<br />

Donovan’s motif counts as given by Anson (Anson counted multiple motifs on a<br />

82<br />

Incised % Total Count<br />

(ds + inc)<br />

Context<br />

40 74 14 26 54 RL2 surface<br />

1647 64 926 36 2573 RL2LayerA<br />

758 66 384 34 1142 RL2LayerB<br />

586 75 199 25 785 RL6LayerA<br />

240 79 65 21 305 RL6LayerB<br />

416 74 143 26 559 SZ8 Surface<br />

789 70 344 30 1133 SZ8levelA<br />

643 66 328 34 971 SZ8levelB<br />

206 67 99 33 305 SZ8levelC<br />

147 67 73 33 220 SZ8levelD<br />

25 71 10 29 35 SZ8levelE


single sherd separately) are the source <strong>of</strong> the data in Table 2 (Anson 1983:175):<br />

<strong>The</strong> low motif count at RF6 is most economically explained as sample-size related,<br />

particularly if the units <strong>of</strong> measurement <strong>of</strong> sample size are reexamined. Decorative<br />

technique abundances are given in Table 3, extracted from various figures in Parker’s<br />

thesis.<br />

Differences in sampling method by site and within sites (Green 1976a: 251-255)<br />

are significant, as Best has pointed out: the RF2 site is the most intensively surface-<br />

collected and Green excavated the largest area here. <strong>The</strong> recovered sample from RF2 is<br />

thus a more complete representation <strong>of</strong> ceramic variability at the site. But even without<br />

making any reference to these data, the ceramic quantities reported by Parker hint at the<br />

same thing. Parker’s rim quantities throw sample sizes for RF6 into stark contrast with<br />

the other two sites. Using sherd count, RF2 has around 12 times as many rims as RF6,<br />

23 times as many using MNI, and 21 times as many counting only those sherds large<br />

enough to measure mouth diameter (providing an unintended size filter). In view <strong>of</strong> this<br />

MNI information on relative sample sizes it seems unlikely that Green’s Jaccard similarity<br />

matrix <strong>of</strong> the three sites is being structured by behavioural variation, and highly likely that<br />

motif presence/absence is strongly structured by sample size. Green’s selection <strong>of</strong> motifs<br />

that were common in some sites to include in the Jaccard similarity matrix is made on the<br />

basis <strong>of</strong> Donovan’s sherd count quantification <strong>of</strong> motif frequency, and is probably an<br />

inadequate compensation for sample-size-related differences in sample richness.<br />

Parker, for vessel forms, like Donovan for motifs, remarks on the similarity <strong>of</strong> the<br />

three sites, also concluding this is “most probably indicating that all belong to a closely<br />

related tradition” (Parker 1981:76). She notes the appearance, though, <strong>of</strong> vessel type 14<br />

(heavily grooved walls) on as many as ten separate vessels at SZ8, absent from other<br />

sites. (<strong>The</strong>se are assumed to be Donovan’s “large, rung-like projections”).<br />

Thus Parker's 1981 conclusion:<br />

83


“It seems that over a period <strong>of</strong> more than a thousand years (1600BC to<br />

400BC) there is in the Western Lapita area little change in the shape <strong>of</strong><br />

pottery vessels, and that we are looking at a very long-lasting and stable<br />

tradition (Parker 1981:115).”<br />

stems more from the earlier slice-in-time view <strong>of</strong> the C14 chronology and the consequent<br />

interpretation <strong>of</strong> Green’s (1978) Jaccard occurrence seriation, than from temporal<br />

behavioural variation in ceramic manufacturing/discard style. An alternative explanation,<br />

where the sites are much closer to each other in age, and potentially contemporaneous or<br />

with overlapping occupation/discard periods, seems a more cautious interpretation <strong>of</strong> the<br />

data, that is, treating Reef-Santa Cruz Lapita as horizon rather than series. This<br />

assessment largely ignores the C14 evidence, which has always tended to suggest that<br />

those RF6 materials which have been dated are younger than dated materials from<br />

SZ8/RF2, and the results <strong>of</strong> current dating research are awaited with interest.<br />

<strong>The</strong> number <strong>of</strong> sites in the Reef-Santa Cruz sample, and their relative occupation<br />

spans are also crucial factors in seriation. Seriation works best when there is a large corpus<br />

<strong>of</strong> sites, and all sites included in the seriation have approximately equal occupation span.<br />

Traditionally this was achieved by rejecting units comprising mixtures from widely different<br />

or excessively long periods, as evidenced by diversity <strong>of</strong> styles. While temporal variation<br />

undoubtedly exists in these three Reef-Santa-Cruz ceramic samples, detailed explication<br />

<strong>of</strong> this variation has yet to be achieved. Best’s call for a re-examination <strong>of</strong> the ceramics<br />

(Best 2002:93) cannot realistically be expected to support a reversal <strong>of</strong> site chronology<br />

as it is unlikely to be possible to construct a robust seriation sequence for Reef-Santa-Cruz<br />

Lapita from a sample <strong>of</strong> only three sites from the Lapita period, four if the Mdailu site is<br />

included (Mccoy & Cleghorn 1988). <strong>The</strong> chances <strong>of</strong> all three being <strong>of</strong> similar duration are<br />

slim, there is a sample-size problem with RF6 at least, and three is simply too few site-<br />

samples to yield a robust seriation. While there are now a number <strong>of</strong> other sites <strong>of</strong> post-<br />

Lapita age recorded in the Reef-Santa Cruz region, these have yet to be seriated, and do<br />

not hold the potential to give a fine-grained chronology <strong>of</strong> Lapita the way<br />

84


a larger sample <strong>of</strong> Lapita sites would. <strong>The</strong> nature <strong>of</strong> this critique needs to be borne in mind<br />

in relation to the Roviana materials from which a seriation is constructed in this thesis: it<br />

suffers from similar problems, with less radiocarbon evidence to test the conclusions.<br />

Difficulties in assessing the relative occupation spans <strong>of</strong> sites are compounded by<br />

differences in the representativeness <strong>of</strong> the ceramic samples between sites and within sites.<br />

Comparison <strong>of</strong> RF2 and RF6 illustrates this. Even the RF2 sample, which must rate as the<br />

most comprehensive and detailed surface collection and area excavation <strong>of</strong> a Lapita site<br />

undertaken to date, is more complete from the southern half <strong>of</strong> the site than the north, due<br />

to more intensive surface collection strategy, a greater degree <strong>of</strong> vertical disturbance to the<br />

south, and concentration <strong>of</strong> excavation in the south. Green felt that the northern area <strong>of</strong><br />

RF 2 was <strong>of</strong> shorter duration than the southern, with deposition <strong>of</strong> ceramics beginning<br />

earlier in the southern area, as evidence by more complex stratigraphy (Green 1976a :255).<br />

This raises a question whether C14 dates from the north <strong>of</strong> RF2, had they been available,<br />

would be similar in age to those from RF6? Certainly some <strong>of</strong> the pottery in the RF2<br />

sample is very similar to some <strong>of</strong> the RF6 pottery, so a simple comparison <strong>of</strong> C14 data<br />

from the RF2 excavations and from RF6 excavations may mislead regarding the stability<br />

<strong>of</strong> the ceramic tradition over time, within the period <strong>of</strong> Lapita production.<br />

Anson’s Early Far Western Lapita:<br />

Anson’s splitting non-classificatory descriptive approach to motif description (Green 1990)<br />

resulted in much sparser data than the Mead/Donovan all<strong>of</strong>orm/motif classification,<br />

potentially creating even greater sampling-related problems for those <strong>of</strong> his analyses that<br />

were motif-occurrence-based. Anson identified nearly 500 different “motifs” in the<br />

analyzed sherd samples, most <strong>of</strong> which occurred only in low frequency, or in a single<br />

instance in many cases. Anson gave “sample sizes” (as motif relative abundances, not as<br />

counts as suggested in Anson 1987) for his sixteen analyzed site assemblages in a lengthy<br />

85


motif appendix, with total motif occurrences (sherd counts?) tabulated by sample in a<br />

subsequent paper (Anson 1986:Table 1). While the overall sherd count <strong>of</strong> motif<br />

occurrences is large for some sites, where sherds-as-vessels analysis has not been<br />

performed we have no way <strong>of</strong> knowing whether these are independent observations. <strong>The</strong><br />

sherd counts and motif occurrences listed by Anson are not necessarily sample sizes in a<br />

behavioural or discard sense.<br />

Using a Robinson coefficient <strong>of</strong> similarity to construct a similarity matrix from<br />

motif relative frequency data for twelve sites from across the Lapita distribution, Anson<br />

found that motifs from Ambitle, Eloaua (Mussau) and Talasea formed a group separate to<br />

the western Lapita sites, in which grouping the Reef-Santa Cruz sites, the Watom sites and<br />

the New Caledonian sites formed separate, and looser sub-groups (Anson 1983:180). On<br />

the basis <strong>of</strong> this grouping and the observations that a similar dichotomy could be seen in<br />

the size and spacing <strong>of</strong> dentate-stamp “teeth”, Anson tentatively defined an “Early Far<br />

Western Lapita”, questioning as he did so whether sample error or other (e.g. functional)<br />

variation could provide a non-temporal explanation for the pattern.<br />

In providing this caveat, Anson sowed the seeds <strong>of</strong> the current critique. <strong>The</strong> spatial<br />

extent <strong>of</strong> the sampling procedure, or the spatial representativeness <strong>of</strong> his samples, is the<br />

first issue. <strong>The</strong> size <strong>of</strong> samples, in terms <strong>of</strong> the number <strong>of</strong> independent behavioural<br />

observations represented, is the second. <strong>The</strong> type <strong>of</strong> behavioural variation that caused<br />

Ambitle, Eloaua and Talasea to cluster separately from the other sites (chrono/stylistic,<br />

functional, geographic) is a third uncertainty.<br />

To compare data from sites in Vanuatu and Fiji for which presence-absence <strong>of</strong><br />

motifs was the only information available, Anson used the MultBET hierarchical cluster<br />

analysis program (he does not give details <strong>of</strong> the similarity coefficient used in this<br />

programme, but it seems likely, like Jaccard’s, to be sensitive to sample size differences in<br />

sparse data, whatever coefficient is used). Kirch has suggested that samples must show<br />

86


diversity saturation (Kintigh 1984) to be comparable (Kirch 1987a:124), but this only<br />

applies to comparing the presence or abundance <strong>of</strong> rare types, while Anson argues the<br />

MultBET clustering results are explicable in terms <strong>of</strong> common types (Anson 1987). This<br />

is supported by similar groupings using the Robinson coefficient <strong>of</strong> similarity with relative<br />

abundance data (Anson 1983:180), and by his sample-size-corrected manually-calculated<br />

motif sharing percentage data (Anson 1987).<br />

As in my critique <strong>of</strong> Green’s 1978 analysis above, the basis on which Anson<br />

discriminated rare motifs from common motifs is questionable, given the impossibility <strong>of</strong><br />

evaluating motif sample sizes in any behavioural sense from sherd counts alone. Similarly,<br />

for the Robinson coefficient results, we cannot know, on the information supplied by<br />

Anson, to what extent the relative motif frequencies as quantified by sherd count are<br />

representative <strong>of</strong> ceramic variability <strong>of</strong> the breakage populations from which they derive.<br />

<strong>The</strong> effects <strong>of</strong> variable occupation span may also be fundamental to the patterning<br />

observed.<br />

<strong>The</strong> implications <strong>of</strong> the spatial/geographic component <strong>of</strong> variation was never fully<br />

discussed. With the exception <strong>of</strong> the three Watom samples, which grouped with the<br />

Western/Eastern Lapita sites, particularly the New Caledonian sites, Anson’s grouping <strong>of</strong><br />

sites largely correlated with geography (Anson 1986:figure 5). <strong>The</strong> similarity between<br />

Watom and western Lapita may be temporal, i.e. Watom and Western/Eastern Lapita is<br />

later than the “far Western” sites; or phyletic ( i.e. either Watom was the origin locality <strong>of</strong><br />

eastwards expansion into the Solomons and Remote Oceania; or Watom represents a back-<br />

migration from the East); or may be reticulate, resulting from some form <strong>of</strong> more<br />

continuous far-west-West-East interaction involving Watom (although the dissimilarity <strong>of</strong><br />

the nearby Ambitle samples to Watom argues against this). Anson in the end plumped for<br />

a temporal explanation, at that time regarding this as a hypothesis in need <strong>of</strong> further<br />

testing.<br />

87


Mussau:<br />

<strong>The</strong> Mussau sites, characterized by Gosden as yielding one <strong>of</strong> the better Lapita ceramic<br />

data-sets <strong>of</strong> the Lapita Homeland Project (Gosden 1991a), cannot be regarded as providing<br />

secure and detailed evidence <strong>of</strong> ceramic change over time until the details <strong>of</strong> the ceramic<br />

analysis are published. Some general review is attempted here, based mostly on preliminary<br />

analyses <strong>of</strong> ceramics (Kirch 1987b, 1988c, 2001). <strong>The</strong> Mussau Lapita Homelands Project<br />

research under Kirch’s leadership located what are claimed to be bedded Lapita deposits<br />

at site ECA area B, documenting a declining frequency <strong>of</strong> dentate-stamping and an<br />

increasing frequency <strong>of</strong> incision with superposition in a stratigraphic/superposition series,<br />

with rim notching reaching a peak in the earlier units (Kirch et al. 1991: Figure 3). <strong>The</strong>re<br />

has recently been a substantial amendment to this preliminary sequence based on C14<br />

evidence, with an early plainware site now understood to predate the dentate Lapita (Kirch<br />

2001:206, 219).<br />

It would be interesting to know whether or not there were sherds from the same<br />

vessels in different zones <strong>of</strong> ECA area B (Summerhayes provides useful information <strong>of</strong><br />

this sort in his Arawe analyses). Zone B1 and B2 were described as highly disturbed by<br />

sand-crab burrowing that was stated to have transported substantial amounts <strong>of</strong> pottery to<br />

the surface (Kirch 2001:86,92), although an argument was made that this had not affected<br />

the waterlogged portion <strong>of</strong> the deposits. Kirch does not theorize the nature and effects <strong>of</strong><br />

bioturbation and swash turbation when the environment <strong>of</strong> deposition was shallow water<br />

in the past. A substantial level <strong>of</strong> burrowing and turbation by marine arthropods,<br />

mammals, flora and fish might be expected to have occurred (Ferrari & Adams 1990);<br />

also in some weather conditions wave processes may have affected the pottery while in<br />

the sea. Regarding the superposition sequence from the upper zones <strong>of</strong> the square,<br />

Gosden’s admonition that sequences need to built from comparable site types (Gosden<br />

1991a) has clearly not been extended to include comparable depositional contexts within<br />

88


an excavation square.<br />

An initial claim for “stratified” evidence from Weisler’s excavation at the EKQ<br />

rockshelter on Elaoua has since been revised. While Kirch now regards the lower<br />

component as “including a late-Lapita component dominated by incised ceramics”(Kirch<br />

2001:214), initial interpretation <strong>of</strong> the ceramic deposit was that it “provided the best<br />

stratified sequence <strong>of</strong> ceramic materials from any Mussau site” (Kirch et al. 1991:151).<br />

Summerhayes, in synthesising the Bismarcks Lapita chronology, incorporated the earlier<br />

view, with a claim for a transition from dentate to incision over time in this deposit<br />

(Summerhayes 2002:27). <strong>The</strong> numerically large sherd sample had an average sherd weight<br />

<strong>of</strong> 1.54 g (Kirch et al. 1991, Weisler 2001), and assessment <strong>of</strong> sample significance is thus<br />

problematic, if even possible. Weisler quantified decorative changes by level using summed<br />

weight by decorative technique, but the total count <strong>of</strong> 26 dentate sherds looks to have been<br />

spread largely at random though the 30 ceramic-bearing levels <strong>of</strong> units one and two (data<br />

for which are illustrated in Weisler 2001:159), and motif analysis has not yet been<br />

published.<br />

<strong>The</strong> “Changing” Face <strong>of</strong> Lapita:<br />

In a departure from either Anson’s Motif inventory approach or Mead’s structural<br />

approach to sherd decoration, Spriggs, reminiscent <strong>of</strong> Specht’s (1968) call for a focus on<br />

whole design, advocated a shift <strong>of</strong> focus to Lapita vessel design in total rather than<br />

quantification <strong>of</strong> decorative motifs represented on sherds (Spriggs 1990). This brings to<br />

mind Shepard’s caution that the brokenness <strong>of</strong> assemblages should not be allowed to<br />

determine units <strong>of</strong> analysis, and that complex pictorial designs are not suited to<br />

element/motif analysis. Spriggs noted that this approach has been severely limited by<br />

generally highly fragmented samples <strong>of</strong> Lapita pottery, but suggested that the Mead<br />

system might not be the most appropriate method <strong>of</strong> analysis for such complex designs,<br />

89


in which he echoed Donovan’s expression <strong>of</strong> the difficulties <strong>of</strong> coding a description <strong>of</strong><br />

complex curvilinear Lapita designs (Donovan 1973:Vol 2: 64, 130-34).<br />

Spriggs suggested that the curvilinear ‘face’ designs formed a coherent<br />

evolutionary chronological series from complex to simple, and that this could be used to<br />

date sites. Spriggs did not discuss the implication that where such a 'series’ was present in<br />

widely-separated sites these would have to have a sufficient level <strong>of</strong> interaction and<br />

convergence in ceramic design (not the same thing as interaction) to have caused parallel<br />

changes in different regions. His analysis is thus founded on an assumed universal Lapita<br />

series rather than derived from detailed regional sequences <strong>of</strong> ceramic change. Some <strong>of</strong> the<br />

principal questions regarding the Lapita ceramic series are to do with whether ceramic<br />

change was synchronous or not across broader regions (Spriggs 2000:355, Summerhayes<br />

2000a:235).<br />

While Mead considered that his structural approach established a phyletic similarity<br />

between all Lapita decoration (as did Green), this is an overstatement for some <strong>of</strong> the<br />

simpler motifs, which could usefully be omitted from analysis as potentially being<br />

analogous similarity rather than homologous. An advantage <strong>of</strong> Spriggs’ approach,<br />

concentrating on the more complex aspects <strong>of</strong> the designs, not stated by Spriggs, is that<br />

potentially analogous similarity <strong>of</strong> some simple designs, widespread throughout the world,<br />

is excluded.<br />

Spriggs’ phyletic approach has recently been elaborated and revised (Ishimura<br />

2002), who, like Spriggs, adopts a view that the overall trend in Lapita decoration, “...the<br />

dynamics <strong>of</strong> chronological change in the design structure....” is one <strong>of</strong> simplification over<br />

time. Mussau is cited as one <strong>of</strong> the regional sequences thought to support this doctrine <strong>of</strong><br />

simplification, which is not wholly congruent with more recent publication <strong>of</strong> Mussau<br />

results which have an early plainware predating Lapita (Kirch 2001:206, 219). While<br />

Ishimura makes no reference to sample-size differences in discussing the presence or<br />

90


absence <strong>of</strong> design variants in site assemblages, his analysis <strong>of</strong> the Santa Cruz sequence,<br />

for example, the occurrence <strong>of</strong> a late variant <strong>of</strong> his “spade” design only in the small RF6<br />

sample but not in the larger RF2 or SZ8 samples is noteworthy. This supports an<br />

interpretation <strong>of</strong> RF6 as including a later date in its occupation span than RF2, unless<br />

absence from RF2 is sampling error.<br />

While Ishimura feels his typological analysis supports a scenario <strong>of</strong> gradual<br />

dispersal, it could also be argued that it supports a scenario <strong>of</strong> rapid dispersal, as some<br />

early variants (Type 3 spade designs for example)are found from Watom to Fiji, showing<br />

no gradual cline, and Type 2 spade designs are only found in the Bismarcks sites, bringing<br />

us back to the uncertainty faced by Anson in trying to interpret essentially the same<br />

dichotomy: is the Type1-Type 2-Type3 evolution <strong>of</strong> spade designs really temporal, or is<br />

this simply geographic variation? Are the Bismarcks designs different because they are<br />

geographically separate or because they are earlier? <strong>The</strong> radiocarbon evidence suggests<br />

the latter, but Type 10 faces occur at Eloaua, RF2 and Site 13 on New Caledonia,<br />

supporting a relatively rapid spread <strong>of</strong> Lapita from West to South within the occupation<br />

span <strong>of</strong> Eloaua, rather than the gradual spread Ishimura suggests. <strong>The</strong> similarity <strong>of</strong> Watom<br />

to New Caledonian Lapita brings to mind processes <strong>of</strong> reticulaton rather than gradual West<br />

to East spread, where potentially Watom is stylistically anomalous in terms <strong>of</strong> its face<br />

designs as a result <strong>of</strong> its geographic location on the stepping-stone route to and from<br />

Remote Oceania. It is not difficult to entertain the possibility <strong>of</strong> a group <strong>of</strong> migrants from<br />

further East becoming established at Watom, for example.<br />

Ishimura demonstrates that a phyletic approach to design seriation is valuable, in<br />

that ideas about the relative occupation spans <strong>of</strong> sites begin to emerge from the lacuna in<br />

which they have languished through the period <strong>of</strong> attribute-based similarity grouping<br />

seriation and heavy reliance on radiocarbon as a primary chronological tool rather than a<br />

confirmatory technique. Ishimura has done us a great favour by bringing focus back to<br />

91


occupation span as a crucial aspect <strong>of</strong> chronology, breaking free from the slice-in-time<br />

view <strong>of</strong> sites so prevalent through the period <strong>of</strong> the New Archaeology. If his phyletic<br />

classification does turn out to be valid, it poses a problem for some <strong>of</strong> the slice-in-time<br />

assumptions, for example a short occupation at RF2 (Sheppard & Green 1991), a site<br />

which along with Eloaua appears on the basis <strong>of</strong> Ishimura’s construct to have a longer<br />

occupation than any other Lapita site from the perspective <strong>of</strong> face designs, while in terms<br />

<strong>of</strong> spade designs it is exceeded in occupation span only by Watom.<br />

While the sampling strategy employed by Green at RF2 is clearly more<br />

comprehensive than the other Lapita samples in Ishimura’s study, which might account for<br />

the richness <strong>of</strong> the RF2 sample, Ishimura’s organization <strong>of</strong> this RF2 material into a phyletic<br />

series spanning Type 3 to Type 14 “face designs” and Type 3 to Type 12 “spade designs”<br />

seems to invoke an alternate explanation, at odds with a slice-in-time view <strong>of</strong> RF2 as a<br />

short occupation.<br />

Whether Ishimura’s Phase 1 (1500-1200BC) is really three centuries in duration,<br />

or whether it is a geographic variant or a slightly earlier variant (perhaps beginning fifty<br />

years earlier?) <strong>of</strong> the more widespread Phase 2 Lapita is not yet known with any certainty.<br />

Ishimura sees his results as supporting Summerhayes’ Early, Middle and Late Lapita, but<br />

one is left wondering how this phyletic analysis gels with Kirch’s recent swing to an early<br />

plainware predating the elaborate designs at Mussau. If his phyletic evolutionary rule from<br />

complex to simple is valid, perhaps the early plainware at Mussau is a functional variant<br />

<strong>of</strong> an even more elaborate set <strong>of</strong> Asiatic designs as yet undiscovered?<br />

A problem with Ishimura’s scheme is that the C14 chronology against which the<br />

phyletic series is tested is <strong>of</strong> greatly different temporal resolution than this phyletic<br />

seriation. <strong>The</strong> C14 data is mostly site-based, while the phyletic data is vessel-based. If ages<br />

<strong>of</strong> “Types” rather than sites in Ishimura’s scheme were available, the prospects for testing<br />

competing theories <strong>of</strong> variability in the ceramic data would be improved.<br />

92


Summerhayes, West New Britain, Anir, and a Three-stage Lapita Ceramic Series:<br />

Summerhayes regarded his west New Britain data as support for Anson’s Early Far-<br />

Western Lapita hypothesis. Like Anson, he saw a temporal dichotomy in the Bismarck<br />

Archipelago data, and extended this to the wider Lapita distribution. Unlike Anson, he<br />

closely identified this dichotomy with form/functional variability within ceramic<br />

assemblages, suggesting that the difference between early and late sites was the loss <strong>of</strong><br />

certain special-function forms, on which particular dentate motifs were found.<br />

His construction <strong>of</strong> a three-phase series for all <strong>of</strong> the Lapita distribution requires<br />

that the New Britain superposition evidence (on which the suggestion that Anson’s theory<br />

was “confirmed” was based) be reviewed in detail. Summerhayes regarded some West<br />

New Britain sites as having lengthy ceramic superposition chronological sequences which<br />

allowed examination <strong>of</strong> the changing nature <strong>of</strong> ceramic production (Summerhayes 2000a:<br />

3-4). He wished to know whether the West New Britain sequence paralleled changes<br />

elsewhere in the Pacific (Summerhayes 2000a:4). Based on his analysis <strong>of</strong> the ceramic<br />

results <strong>of</strong> several West-New Britain excavation conducted during the Lapita Homeland<br />

Project, followed by wider comparisons, he concluded that the stylistic province models<br />

<strong>of</strong> Green (Green 1978, 1979) and Kirch (Kirch 1997) could be replaced with a simple three<br />

stage overall sequence <strong>of</strong> ceramic change, where the far-western, western and eastern<br />

styles become early, middle and late Lapita (Summerhayes 2000a:235). Sites FNY and FOJ<br />

were regarded by Summerhayes as having long sequences, with heavy use made also <strong>of</strong> site<br />

FOH stratigraphy (squares D, E, and F) also, in defining a temporal directional change in<br />

pottery style. Summerhayes’ evidence from these sites will be reviewed in order to evaluate<br />

his Lapita Ceramic Series.<br />

Site FOH:<br />

FOH was located on a sand spit, with the dense pottery concentration below the water<br />

93


table (as at site ECA in the Mussau group). Excavation over four field seasons included<br />

geomorphological investigation (Gosden & Webb 1994, Summerhayes 2000a:21-22). <strong>The</strong><br />

lower levels <strong>of</strong> the site were thought to have originated as the discard from stilt villages<br />

into shallow water, and these lower 45-50cm <strong>of</strong> deposit were excavated in spits forming<br />

Summerhayes’ analytical units A-E for squares D, E and F, with A being the basal spit and<br />

E the uppermost, the latter in the partly-concreted sand layer between the brackish<br />

groundwater and the salt tidal incursion.<br />

Summerhayes reported that sherd counts decreased dramatically up through these<br />

five levels (Summerhayes 2000a:43), and vessel refitting found a high degree <strong>of</strong> vessel<br />

completeness (see illustrations Summerhayes 2000a:67-70) with many joins between these<br />

levels. Summerhayes considered that as only three co-joins from level E were with<br />

underlying units there was little evidence for vertical disturbance (Summerhayes 2000a:22-<br />

23) but I would argue that 100% <strong>of</strong> the joins in level E were to other layers, and that<br />

Summerhayes' conclusion that little disturbance had occurred since deposition is not well<br />

supported by his data. <strong>The</strong> number <strong>of</strong> co-joins to other levels for each level (Summerhayes<br />

2000a: Table 3.1 on p23) seem consistent with the analyzed sample size for each layer<br />

(Summerhayes 2000a: p44 Table 5.1) and a reasonable explanation for the overall pattern<br />

<strong>of</strong> co-joining is that turbation <strong>of</strong> some description has occurred across the four lower<br />

excavation levels. Summerhayes’ claim that pottery assemblages from these levels can be<br />

regarded with confidence as superposed temporal sets is open to question on this basis<br />

alone.<br />

Are there perhaps taphonomic or other sampling explanations for the changes in<br />

decorative technique and motif observed through the levels? <strong>The</strong>re is clearly a different<br />

chemical weathering regime at work in Layer E as evidenced by the presence <strong>of</strong> the<br />

concretion here, and the decorative and stylistic temporal changes suggested by<br />

Summerhayes are furthermore evidenced by a sample <strong>of</strong> only 49 analyzed sherds from this<br />

94


upper concretion level, including18 decorated sherds; thus sample size effects might also<br />

easily account for the “temporal” patterning noted for this part <strong>of</strong> site FOH. This is<br />

particularly so where vessel completeness is this high, as evidenced by Summerhayes'<br />

excellent illustrations <strong>of</strong> vessel families, with many sherds potentially from single vessels.<br />

<strong>The</strong> vessel sample may consequently be much smaller than the sherd count, which may not<br />

provide independent observations in these circumstances.<br />

FOH squares G1 and G2 were situated 15m south <strong>of</strong> squares D, E and F discussed<br />

above, and yielded 2883 sherds, <strong>of</strong> which 156 were analyzed. Stratigraphy seems almost<br />

identical to squares D, E, and F, with layer three and layer four seeming likely to equate<br />

to the previous units A-D <strong>of</strong> squares D, E and F (Summerhayes 2000a:23). Layer 3 yielded<br />

a total <strong>of</strong> eight decorated sherds, while Layer 4 yielded 89 (Summerhayes 2000a:91-92).<br />

Summerhayes argues in Chapter 10 that the assemblage from this square is later than the<br />

assemblage from Squares D, E and F, units A-D, and contemporaneous with unit E (the<br />

sample <strong>of</strong> 18 decorated sherds). As stated above, a chrono-stratigraphic distinction<br />

between any <strong>of</strong> these lower units seems unwarranted on the laudably detailed evidence<br />

presented by Summerhayes, and a stylistic distinction faces the serious and fundamental<br />

questions <strong>of</strong> whether one is seeing activity/functional patterning across a single occupation<br />

unit or chronological change between areas, or whether this is simply sampling error. <strong>The</strong><br />

differences in the frequency <strong>of</strong> dentate-stamp decoration may also relate to<br />

(contemporaneous?) functional differences in vessel form (Summerhayes 2000a:101) rather<br />

than differences in time.<br />

Site FOJ:<br />

Summerhayes’ evidence for stratified stylistic change at FOJ (one line <strong>of</strong> evidence for his<br />

contention that the direction <strong>of</strong> ceramic change is known) is examined below. Units B, C<br />

and A seem broadly uniform in their characteristics, the major difference being that unit<br />

95


A has about six dentate-stamped sherds less than would be expected from the sample size.<br />

In view <strong>of</strong> the difference between these depositional contexts (B/C, D 1n shallow water<br />

and A on land) attributing such differences to temporal change in ceramic production is<br />

problematic.<br />

Gosden and Webb describe unit A as follows:<br />

“<strong>The</strong> bulk <strong>of</strong> sediment accumulation probably occurred during storm<br />

activity, when sea-level is probably elevated.... (Gosden & Webb<br />

1994:41)”, and comprising: “...a unit <strong>of</strong> light coloured sand...roughly 1m<br />

thick...contains relatively small amounts <strong>of</strong> pottery and obsidian, but<br />

moderate amounts <strong>of</strong> shell....Occasional mumu (oven) stones are also<br />

present....<strong>The</strong> reduced percentage <strong>of</strong> artefactual material in this deposit<br />

indicates that it is unlikely that there was still a village located directly on<br />

the site (Gosden & Webb 1994:40-41).”<br />

<strong>The</strong> presence <strong>of</strong> the oven stones was for unspecified reasons considered evidence for<br />

terrestrial use <strong>of</strong> this location (presumably because these were unlikely to remain in<br />

suspension in the water column, but transport by rolling onto the beach ridge as a result<br />

<strong>of</strong> storm-wave action was apparently not considered). Summerhayes states that: “It is<br />

within this layer that occupation occurred on dry land.... (Summerhayes 2000a:24)”,<br />

implying the ceramics from this layer were from a settlement on dry land. Many <strong>of</strong> the<br />

cultural materials in this layer, including pottery, might derive from material originally<br />

deposited in the sea at an earlier time (contemporaneous with the lower units?) and<br />

redeposited by storm waves to form a beach ridge.<br />

Site FNY:<br />

Summerhayes’ analysis <strong>of</strong> the FNY ceramics concluded that dentate-stamp decoration<br />

drops in frequency in the upper <strong>of</strong> three units, and noted trends in vessel form from lower<br />

to top unit (Summerhayes 2000a:125-126). In Chapter 10, this is re-stated as:<br />

“incised and fingernail impressed decoration increasing in the upper units,<br />

although here dentate decoration is always dominant. (151)"<br />

Gosden and Webb interpret the lower two units, both brown clay separated by a white<br />

96


sand lens, as back-swamp deposits during a period <strong>of</strong> Lapita occupation and beach-ridge<br />

formation (Gosden & Webb 1994:46-47, Summerhayes 2000a:25). <strong>The</strong> top unit comprised<br />

black shell midden with Lapita and Late-prehistoric (my emphasis added) pottery mixed<br />

(Gosden & Webb 1994:47). Summerhayes describes this top unit as comprising<br />

“A top black midden layer (1.2m) incorporating post Lapita (my emphasis<br />

added) and Lapita remains (Summerhayes 2000a:25).”<br />

Summerhayes’ use <strong>of</strong> the term “post Lapita”, suggests episodes <strong>of</strong> ceramic<br />

deposition contributing to the upper unit are not widely separated in time. <strong>The</strong> zone<br />

between the upper mixed unit and the lower, purely Lapita units yielded several<br />

radiocarbon determinations which varied between 1290 and 690 cal BP, with a Lapita-age<br />

estimate from an oyster shell lower down the brown clay. (Summerhayes 2000:25). It<br />

seems well attested from the C14 evidence and from the Lapita and late-prehistoric pottery<br />

styles represented that the upper unit represents a mixture <strong>of</strong> materials from widely<br />

different periods, while at least the upper portions <strong>of</strong> the brown clay also comprise mixed<br />

deposits. <strong>The</strong> absence <strong>of</strong> recent styles in the brown clay lower units, and a single Lapita-<br />

age date, do not provide evidence that the lower clay is temporally un-mixed, merely that<br />

the clay formed during a period in which only Lapita age materials were available as either<br />

a natural or cultural sediment supply. If this site comprises variable mixtures <strong>of</strong> styles from<br />

widely separated periods (c.1000 BC Lapita and c.1000 AD late-prehistoric) it should not<br />

be regarded as a superposition sequence <strong>of</strong> change in the Lapita ceramic complex over<br />

time (see for example Summerhayes 2000a:151), and the status <strong>of</strong> this unit in subsequent<br />

seriations (for example Summerhayes 2000a:152 Figure 10.1) should be interpreted with<br />

this in mind.<br />

Summerhayes’ construction <strong>of</strong> the direction <strong>of</strong> ceramic change does not<br />

convincingly link superposition assemblage variability to time, by under-utilizing Gosden<br />

97


and Webb’s study <strong>of</strong> site formation processes, and by comparisons between samples <strong>of</strong><br />

dissimilar depositional type (see Phelan 1997:36-38). Summerhayes’ conclusion that<br />

“the Arawe assemblages largely agree with other Lapita assemblages which<br />

show both a decrease in dentate and an increase in linear incision over<br />

time....”<br />

and that the Arawe assemblages “confirm” the trends noted by others in this regard<br />

(Summerhayes 2000a:151) is not well supported by the evidence presented. Summerhayes<br />

went on to use this directional theory in a series <strong>of</strong> evolutionary seriations, using<br />

decorative technique frequency and motif occurrence.<br />

Summerhayes’ Seriation Using Decorative Technique/ Vessel Form:<br />

Having concluded that he had confirmed the temporal direction <strong>of</strong> ceramic change,<br />

Summerhayes constructed a seriation <strong>of</strong> West New Britain assemblages based on<br />

decorative technique frequency (Summerhayes 2000a:152) Here excavation squares are<br />

treated as single-period samples rather than time-transgressive units as in the previous<br />

stratigraphic analysis, with all layers being included as the site sample. <strong>The</strong>re is no<br />

discussion in relation to the seriation <strong>of</strong> the relative occupation spans <strong>of</strong> these units, or the<br />

degree <strong>of</strong> temporal mixing <strong>of</strong> separate periods, which are both important aspects <strong>of</strong><br />

seriation method as outlined in Chapter 1.<br />

Decorative technique and vessel form were found to be highly correlated. Sites<br />

failed to seriate when controlled for vessel form in his decorative technique seriation<br />

(Summerhayes 2000a:p153 Figure 10.5), leading Summerhayes to conclude that what<br />

changed over time was that some vessel forms dropped out. Alternative explanations for<br />

the failure <strong>of</strong> these sites to seriate when controlled for vessel form is that the classificatory<br />

units <strong>of</strong> decorative technique used were too coarse, or not salient to stylistic variation, or,<br />

more serious, that stylistic change did not occur across his “sequence”. This last<br />

possibility contrasts with Dunnell’s definition <strong>of</strong> style as evolutionary drift. Some form<br />

98


<strong>of</strong> stylistic change must occur within a single vessel form, even in the most conservative<br />

milieu, unless extraordinary measures are taken to prevent change. <strong>The</strong> difficulty is in<br />

identifying that change. It is possible therefore that the “seriation” is simply a similarity<br />

grouping <strong>of</strong> functional variability, that has little or nothing to do with time. <strong>The</strong> most<br />

striking example <strong>of</strong> this is that his seriation separates FOH squares DE&F from FOH<br />

square G, despite convincing stratigraphic similarities between the lower levels <strong>of</strong> these<br />

excavation units and a spatial separation <strong>of</strong> only 15m.<br />

Summerhayes’ Lapita Motif Analysis and Early/Middle/Late Lapita:<br />

Summerhayes went on to construct a motif presence/absence/sharing similarity grouping<br />

for a range <strong>of</strong> Lapita sites from across the Lapita region, using Anson’s motif inventory<br />

as units <strong>of</strong> classification (Summerhayes 2000a:table 10.5). It is important to note that<br />

presence/absence <strong>of</strong> motifs rather than motif frequencies were the base data for his<br />

analyses. Anson’s splitting classification creates a large inventory <strong>of</strong> motifs for which<br />

there is sparse data. Motif sharing may be quite arbitrary in these circumstances (Kirch<br />

1987a).<br />

Summerhayes performed a cluster analysis <strong>of</strong> motif presence/absence, and a<br />

Principal Components Analysis <strong>of</strong> the number <strong>of</strong> shared motifs between sites. He used<br />

Ward’s clustering method (which tends to force outliers into groups and produce tight<br />

clusters in the data or different clusters compared to some other methods such as average<br />

linkage), but does not state what sort <strong>of</strong> similarity information was input, whether number<br />

<strong>of</strong> shared motifs, or some similarity matrix such as Jaccard. This clustering method has<br />

been widely used in archaeology for numeric data such as trace-element concentrations<br />

(Shennan 1997:244) and produced what I see as anomalous groupings with Anson’s motif<br />

presence-absence data. <strong>The</strong> Reef-Santa Cruz sites formed a cluster separate from all other<br />

sites, and Summerhayes states this is a sample-size effect, without discussion <strong>of</strong> the<br />

99


independence <strong>of</strong> observations and how sample size is measured. As shown above, the<br />

Reef/Santa Cruz Lapita sites include quite large samples from RF2 and SZ8, and also a<br />

very small sample, RF6, so sample size should not cause these to cluster separately.<br />

A PCA plot <strong>of</strong> the first and third PCA components agreed with the cluster analysis,<br />

in separating the Bismarcks sites into similar groupings to those obtained in his vessel-<br />

form/decorative technique analysis. As in his decorative technique frequency seriation,<br />

FOH squares fall into separate clusters, Square G being more similar to Yanuca in Fiji<br />

than to squares D, E and F 15 metres away. Summerhayes is not clear about the type and<br />

structure <strong>of</strong> his primary data: he refers to performing “multivariate analysis on the<br />

presence/absence <strong>of</strong> motifs in all sites” (Summerhayes 2000a: 160) but also captions<br />

Figure 10.2 (his PCA plot) “Motif Sharing”, and does not provide the raw data input to<br />

his analyses, rendering his multivariate computer analysis something <strong>of</strong> a black box. PCA<br />

is most appropriate for numeric data (Shennan 1997:298), and Summerhayes may have<br />

done better to use sherd counts or MNV to identify a selection <strong>of</strong> common motifs with<br />

Correspondence Analysis used as a seriation technique on this selection.<br />

Comparisons are made (Summerhayes 2000a: Table 10.5, Fig 10.12, Table 10.7)<br />

without discussion <strong>of</strong> the occupation spans <strong>of</strong> sites or potential functional variability<br />

within sites. If site occupation spans are lengthy or variable, and function varies<br />

substantially across space within sites, assemblages may comprise an inextricable mix <strong>of</strong><br />

temporal and functional information, without any clear temporal patterning identifiable by<br />

PCA or cluster analysis.<br />

Summerhayes’s statement that “...differences in (sample) size are seen in the first<br />

PCA component...(Summerhayes 2000a:160, 161)” suggests that it is indeed motif-<br />

sharing counts being input to the PCA. Eigenvalues <strong>of</strong> the PCA components are not given,<br />

and he chooses to plot the third component against the first, without discussion <strong>of</strong> the<br />

significance <strong>of</strong> either the second component or the third. What is the significance <strong>of</strong> the<br />

100


second hidden component, which must account for more variation in the data than the<br />

third? Furthermore, if the first component is primarily a reflection <strong>of</strong> sample size, why use<br />

it in the plot, or ascribe temporal significance to the clusters thus formed?<br />

<strong>The</strong> results <strong>of</strong> one further analysis, using Anson’s manual motif-sharing similarity<br />

method (Summerhayes 2000a :160-163 and table 10.7), contrast sharply with the previous<br />

multivariate numerical taxonomy results in some respects. FOH square G, more similar to<br />

Yanuca in Fiji than to the adjacent excavation <strong>of</strong> FOH squares D, E, F in his PCA and<br />

cluster results, appears, using the Anson manual method, to be more similar to Squares D,<br />

E and F than to anywhere else, as might be expected from the same layer <strong>of</strong> the same site,<br />

separated by 15m <strong>of</strong> unexcavated ground.<br />

Summerhayes goes on to use these motif analysis results to bolster his argument,<br />

following Anson, that there is a dichotomy in the Lapita assemblages <strong>of</strong> west New Britain,<br />

and that this is present also in the wider Lapita universe, ascribing a temporal dimension<br />

to this dichotomy, crosscutting a regional (spatial) dimension <strong>of</strong> variation.<br />

“Similar changes in the Lapita decorative system occur in the west and<br />

east....the product <strong>of</strong> information exchange which necessitates the<br />

movement <strong>of</strong> people...” (Summerhayes 2000a:233).<br />

In his discussion <strong>of</strong> the motif results, he is clear that the similarities noted between<br />

assemblages relate to the presence or absence <strong>of</strong> particular vessel forms in the recovered<br />

samples, but still does not talk about the extent to which contemporaneous within-site<br />

functional variation, rather than temporal variation, might account for his groupings <strong>of</strong><br />

West New Britain (and other) Lapita assemblages.<br />

While the inventory <strong>of</strong> vessel functions in use in Lapita communities may have<br />

shifted over time, it is also entirely possible that vessel forms/functions were unevenly<br />

distributed across such communities, and that telephone-booth archaeology is hitting<br />

varying mixes <strong>of</strong> the plain stuff and the fancy stuff. FOH in the Arawes may be the spatial<br />

lesson in this regard, as Sheppard and Green have suggested also for RF2 in the Southeast<br />

101


Solomon Islands (Sheppard & Green 1991).<br />

My conclusion is that the Early/Middle/Late Lapita construct is an example <strong>of</strong><br />

Dunnell’s “...inseparable hodgepodge...” <strong>of</strong> geographic, functional and temporal variation<br />

(Dunnell 1970:310), to which could be added uncertainties regarding taphonomic bias,<br />

quantification <strong>of</strong> motifs and the degree to which samples are representative.<br />

Early/Middle Lapita in the Anir Group:<br />

Recent extension <strong>of</strong> the west New Britain research through a program <strong>of</strong> survey and<br />

excavation in the Anir group has augmented the corpus <strong>of</strong> Bismarck Archipelago Lapita<br />

samples, with the most notable assemblages excavated from a single productive testpit at<br />

Malekolon and from a grid <strong>of</strong> 14 productive testpits at Kamgot (Summerhayes 2000b,<br />

2001, 2002). Summerhayes suggests Malekolom TP4 (where the main cultural layer was<br />

located) was deposited on land, near to the then shoreline, while the extensive deposit at<br />

Kamgot was thought to have been deposited either in the water or at the water’s edge<br />

(Summerhayes 2000b). Summerhayes slotted Malekolon and Kamgot sample assemblages<br />

into his early-middle-late Lapita chronology, using the same methods <strong>of</strong> vessel form<br />

classification, frequency <strong>of</strong> dentate-stamping (usefully quantified using both sherd count<br />

and MNI this time) motif-based assemblage grouping, and obsidian source frequency).<br />

Summerhayes' use <strong>of</strong> C14 evidence to test his Early/Middle temporal construct<br />

(Summerhayes 2002) is not an independent confirmation <strong>of</strong> the temporality <strong>of</strong> variation.<br />

While a plot <strong>of</strong> some <strong>of</strong> these dates appears to be convincing evidence <strong>of</strong> the temporal<br />

primacy <strong>of</strong> Summerhayes’ “early Lapita”, his process <strong>of</strong> chronometric hygiene excludes<br />

any dates which don’t agree with his temporal construction. Thus Beta 55323, a date <strong>of</strong><br />

2880 ±70BP from Adwe squares DEF (“early” Lapita) is rejected solely because it does<br />

not fit his temporal hypothesis, and similarly, WK7564, a date <strong>of</strong> 2960-2760 cal BP for the<br />

Kamgot “early” Lapita deposit is rejected “because it is statistically distinguishable<br />

102


from the other four dates (from that context)”.<br />

Summerhayes also rejects Kamgot dates that have a wide confidence interval<br />

(ANU 11193, ANU 11190, ANU 11191). One <strong>of</strong> these dates(ANU 11190: 2750-1530 cal<br />

BP), rather than limiting “the ability to make fine discriminations in the chronology”would<br />

tend to undermine his chronological hypothesis (Kamgot as early Lapita) in spite <strong>of</strong> its<br />

broad confidence interval. Summerhayes’ conclusion that “<strong>The</strong>se new determinations<br />

confirm the chronology for sites in that region that were based on regional comparisons<br />

<strong>of</strong> pottery decoration, form and production.” is thus not unequivocally supported by<br />

current data. Any evidence for occupation span (non-overlapping ages) is ruled out rather<br />

than put to good use to understand what sort <strong>of</strong> temporal mixture the samples comprise.<br />

<strong>The</strong> Watom Lapita Series:<br />

Watom, SAC location, had Lapita deposits preserved with a minimum <strong>of</strong> recent<br />

disturbance under a volcanic ash fall (Green & Anson 2000b:37). Anson concluded that<br />

the stratified deposits containing pottery at SAC locality “enabled the chronological<br />

position <strong>of</strong> unstratified material from SAD and museum collections to be inferred (Anson<br />

2000)”. Best critiqued Green and Anson's temporal conclusions, that Watom Lapita dated<br />

500BC-AD200, suggesting the younger C14 dates obtained postdated the Lapita pottery<br />

recovered (Best 2002:87).<br />

Green and Anson noted that initial Lapita occupation at Watom SAC zone C2<br />

predates their earliest C14 date (Green & Anson 2000b:38), preferring to discard a<br />

Trochus shell date <strong>of</strong> 3490±80BP (calibrated at 1-sigma to 1509-1350BC) in the surface<br />

<strong>of</strong> Zone D as being unrelated to Zone C2 occupations (Green & Anson 2000b:39). Green<br />

and Anson present ample and explicit evidence that the buried Zone C1 paleosol is a<br />

loamy gardening soil incorporating cultural materials (Green & Anson 2000b:42), but this<br />

103


stratigraphy is elsewhere described as “...two cultural occupation layers, C1 and C2....<br />

(Green & Anson 2000b: 35).” Zone C1 was dated using several partly dissolved Tridacna<br />

maxima shells from the base <strong>of</strong> the layer (Green & Anson 2000b:38). C2 was dated using<br />

a stack <strong>of</strong> Tridacna shells in a pit feature unconformably overlain by zone C1. <strong>The</strong>se were<br />

calibrated and interpreted to yield a time span <strong>of</strong> 400BC to about AD100. As Green and<br />

Anson are careful to point out, features associated with Zone C1 could only be identified<br />

as intrusions into the surface <strong>of</strong> C2, confirming a turbation unconformity as the formation<br />

process <strong>of</strong> Zone C1, as do the smaller and more fragmented bones and potsherds from<br />

Zone C1 in comparison to Zone C2.<br />

For Zone C2 it was possible to demonstrate sequences <strong>of</strong> superimposed features.<br />

Fom the section drawings, it looks as though a burial ground was cut in from C1 level,<br />

prior to the formation <strong>of</strong> the C1 unconformity, and which postdates some pottery and<br />

obsidian from C2 (Green & Anson 2000b:45). In view <strong>of</strong> these formation processes,<br />

Anson’s conclusion that “A fair degree <strong>of</strong> motif sharing between layers C1 and C2 at SAC<br />

gives an overall impression <strong>of</strong> continuity, with no evidence <strong>of</strong> a dramatic break....(Anson<br />

2000:133)” ignores the fact that zone C1 clearly incorporates materials mixed in from, and<br />

originating as the upper levels <strong>of</strong> part <strong>of</strong> Zone C2, during the creation <strong>of</strong> the C1 turbation<br />

unconformity. Furthermore, a complex history <strong>of</strong> excavated unconformities is evidenced<br />

in Zone C2 stratigraphy (Green & Anson 2000b:44), which must have acted to bring<br />

materials from the earliest facies <strong>of</strong> C2 into younger stratigraphic features, and into<br />

stratigraphic association with Mopir obsidian, thought to be unavailable in Lapita times<br />

(White & Harris 1997). <strong>The</strong> burial pits cannot have been “...sealed in by layer C1 (Green<br />

& Anson 2000b:45)." given the formation processes so well documented for C1. Materials<br />

constituting C1 must in part derive from the upper levels <strong>of</strong> what was originally C2 prior<br />

to gardening.<br />

Rather than there being no evidence for a dramatic break, an important question<br />

104


is whether there is historical continuity <strong>of</strong> occupation beneath the ash from the time <strong>of</strong><br />

initial Lapita deposition to the time <strong>of</strong> the use <strong>of</strong> the site as a burial ground. As Best points<br />

out, some <strong>of</strong> the zone C1 pottery is not necessarily from a late-Lapita context. <strong>The</strong>re is no<br />

evidence to rule out the possibility that some <strong>of</strong> this pottery could substantially post-date<br />

the Lapita-horizon, while pre-dating the Zone B ash, with an intervening hiatus <strong>of</strong> several<br />

centuries, while those sherds that are clearly Lapita may be incorporated by the well-<br />

evidenced mixing which was the origin <strong>of</strong> C1. How does one tell whether Lapita pottery<br />

in a deposit is mixed with post-Lapita, when post-Lapita sequences are so ill defined as in<br />

the Bismarck Archipelago? Given the clear evidence for excavation into C2 after<br />

deposition <strong>of</strong> Lapita pottery, which involved burial practices not known elsewhere in Near<br />

Oceania for Lapita, it seems reasonable to worry that the stratified sequence <strong>of</strong> “change”<br />

may be blurred by these formation processes. Anson’s analysis <strong>of</strong> ceramic style by layers<br />

is critiqued below in view <strong>of</strong> these formation processes and uncertainties regarding<br />

occupation spans and historical continuity <strong>of</strong> deposition.<br />

Nail-impressed sherds were found in layer C1 with Lapita, but not in Layer C2.<br />

This may simply be a sample size effect (Green & Anson 2000b:77), or a mixing <strong>of</strong><br />

different periods.<br />

Anson found that Lapita motifs from SAC C2 were more similar to SAD and<br />

Meyer’s collection than to SAC C1, although SAD and the Meyer collection are more<br />

similar to each other than to C2 (Anson 2000:132). Anson concludes that SAC Zones C1<br />

and C2 represent temporal stratigraphy, based on a relatively low degree <strong>of</strong> motif sharing.<br />

Given that the 38 motif occurrences listed for “SAC L1" and “SAC L2" in Table 1 are<br />

represented by sparse frequency data, in 29 instances by single sherds per layer, and never<br />

by more than two sherds per layer for the remaining nine sherds, it is not surprising that<br />

these layers have a low degree <strong>of</strong> motif sharing. Contrast this with the combined Meyer<br />

sample where single motifs are represented by up to 16 sherds per layer, (motif 429) and<br />

105


it seems clear that there are major unresolved sampling issues here which have not been<br />

discussed.<br />

Anson, in grouping collections using motif sharing, is attempting to produce an<br />

occurrence seriation, without due regard to the effects <strong>of</strong> sample sizes on motif<br />

occurrences, and also without consideration <strong>of</strong> what his counts mean behaviourally, and<br />

how sample size might be quantified. Even assuming that his method <strong>of</strong> quantification is<br />

salient to the completeness properties <strong>of</strong> the samples, if seven out <strong>of</strong> 16 motif occurrence<br />

at SAC C2 are shared with SAD, and four out <strong>of</strong> 18 motif occurrences in SAC C1 are<br />

shared with SAD, it cannot be concluded with confidence that there is a significant<br />

difference. <strong>The</strong> difference <strong>of</strong> three motifs may simply be due to chance.<br />

Anson’s analysis <strong>of</strong> the complete corpus <strong>of</strong> Watom pottery states that most <strong>of</strong> the<br />

Mussau Lapita sequence is much earlier than Watom Lapita on radiocarbon evidence<br />

(Anson 2000:120, footnotes), but Green and Anson noted that initial Lapita occupation<br />

at Watom SAC zone C2 predates their earliest C14 date (Green & Anson 2000b:38). In<br />

a similar vein as for the Reef-Santa Cruz sites, I suggest that insufficient consideration has<br />

been given to the possibility that the SAC site and the Reber/Rakival area generally have<br />

been the focus <strong>of</strong> Lapita occupation for an extended period, potentially beginning many<br />

centuries earlier than 400BC, and that individual site/excavation unit assemblages may be<br />

palimpsests <strong>of</strong> a variety <strong>of</strong> occupational events and pottery styles within the Lapita period,<br />

with lengthy hiatuses in occupation not ruled out by the evidence. Ishimura’s phyletic<br />

analysis would tend to support this alternate view, given the diversity <strong>of</strong> face and spade<br />

designs from Watom (Ishimura 2002).<br />

<strong>The</strong> pottery styles illustrated (e.g. Green & Anson 2000a: Figure 1) seem to me to<br />

have equivalents in all three analyzed Reef-Santa Cruz assemblages and also in Mussau<br />

and the Arawes. <strong>The</strong> possiblity that initial domestic occupation began substantially earlier<br />

than 400BC, potentially as early as the dated Trochus shell from Zone D (which is<br />

106


consistent with the earlier dates from house-posts at Mussau) is not precluded by Green<br />

and Anson’s detailed excavation data and materials analyses. <strong>The</strong> large numbers <strong>of</strong> Lapita<br />

sherds recovered from Zone C1 are from a chronostraphic unit containing ceramics<br />

potentially manufactured between c. 1400BC and c. 100BC, whereas some <strong>of</strong> the earlier<br />

stratigraphic units in Zone C2 incorporate materials manufactured over an earlier, shorter<br />

period.<br />

Wahome’s Seriation <strong>of</strong> Admiralties Pottery Assemblages:<br />

Wahome’s comparison <strong>of</strong> incised/applied pottery site across Island Melanesia has been<br />

called into question by new data from Vanuatu and by criticism <strong>of</strong> classificatory units used<br />

as being too coarse to be salient to the research question (Bedford 1999:167-189). For the<br />

Admiralties, Wahome came up with a local sequence <strong>of</strong> four periods, from Lapita, through<br />

early post Lapita, late post-Lapita and late prehistoric (Wahome 1999:112-118). This<br />

series conflicts with Ambrose’s assessment <strong>of</strong> historical continuity <strong>of</strong> the Admiralties<br />

sample. According to Ambrose a gap <strong>of</strong> up to seven centuries separates Lapita pottery<br />

from subsequent styles (Ambrose 1991, Bedford 2000:183).<br />

Potential problems with historical continuity aside, Wahome’s approach is<br />

interesting for his use <strong>of</strong> Correspondence Analysis. Wahome used Correspondence<br />

Analysis to seriate fourteen Admiralty Islands pottery samples (Wahome 1999:33-34 and<br />

96-99). Both Lapita-style and later/other pottery variants were included in his attribute and<br />

attribute-combination analyses. Wahome was concerned to identify attributes and attribute<br />

combinations that varied over time. Wahome saw the summary variables in his analysis as<br />

separating functional and temporal variation. He looked also for “reliable sequences”<br />

in these dimensions <strong>of</strong> variability by cross checking against stratigraphic trends in<br />

excavated sites. He had difficulty seriating some units and concluded that some <strong>of</strong> the<br />

division <strong>of</strong> excavated sites into stratigraphic/superposition units was unjustified, due to<br />

107


lack <strong>of</strong> variability between units (Wahome 1999: section 5.3). His seriation became more<br />

coherent when these units were lumped.<br />

His Lapita unit remained distinct from all other units in his analysis (which is<br />

consistent with Ambrose’s suggestion <strong>of</strong> a lack <strong>of</strong> historical continuity). <strong>The</strong> site <strong>of</strong> Mouk<br />

yielded both Lapita and post-Lapita evidence (Wahome 1999:5.3, 5.5.1). Three “early<br />

post-Lapita” sites dating to around 2000BP could be seriated using the frequency <strong>of</strong> shell<br />

impression, with shell impression absent earlier than 1900BP (Wahome 1999:5.4.1,<br />

5.5.2.1). A group <strong>of</strong> ten “Late post-Lapita” sites seriated in the first dimension (66.8%<br />

inertia) on the frequency <strong>of</strong> cross-hatch incisions, rolled-rims, shell impressions, and<br />

combinations <strong>of</strong> these. A second dimension <strong>of</strong> variability (17%) comprised the frequency<br />

<strong>of</strong> deep notching, flat lips on rolled rims, and notching on the plain lips <strong>of</strong> incurving rims.<br />

<strong>The</strong> first dimension was ascribed to time by comparison with stratigraphic relationships<br />

between units, and the second was disregarded. Eight <strong>of</strong> the ten late post-Lapita units<br />

included fingernail and/or fingertip impressions, with one site showing a greater diversity<br />

<strong>of</strong> motifs in these decorative techniques. A “highly diversified range” <strong>of</strong> incised decorations<br />

was recorded in all these units, and punctations were present in seven sites.<br />

His final grouping <strong>of</strong> 19 “Late prehistoric” sites (post-800BP) was differentiated<br />

primarily (inertia <strong>of</strong> 24%) on the frequency <strong>of</strong> brushing, in which respect one site in the<br />

group differed significantly from the others. A number <strong>of</strong> attributes/combinations<br />

contributed to the second dimension (inertia was eighteen percent). He considered that<br />

most <strong>of</strong> this variability correlated with space rather than time or function. <strong>The</strong>re was a<br />

wider range <strong>of</strong> vessel forms in evidence than in the preceding period, with brushing and<br />

perforation more common (Wahome 1999:5.4.3, 5.5.2.3).<br />

Overall, taking into account units <strong>of</strong> classification used, and the use <strong>of</strong> seriation<br />

in combination with stratigraphy and surface sites, Wahome’s approach is similar in<br />

method to the Irwin/Frost school, but with the more up-to-date technique <strong>of</strong><br />

108


correspondence analysis substituted for similarity matrix assemblage grouping. Like Irwin,<br />

Wahome put some effort into identifying temporal variation, primarily through comparison<br />

with stratified superposition chronologies. Wahome’s Lapita and Early post Lapita phases<br />

cannot be considered to form a historically continuous series though, as historical<br />

continuity is poorly evidenced at best, and Wahome makes no argument in support <strong>of</strong><br />

heritable continuity. Thus Lapita remains a horizon rather than a series in the Admiralties.<br />

<strong>The</strong> Buka sequence, to the north <strong>of</strong> Bougainville, was established in its current<br />

incarnation primarily through the pioneering work <strong>of</strong> Specht (Specht 1969), and a further<br />

indirect contribution came out <strong>of</strong> the Lapita Homeland Project on Nissan (Spriggs 1991).<br />

More recently, the Buka area has been substantially researched by Wickler (Wickler<br />

2001).<br />

Specht on Buka:<br />

Specht constructed a detailed sequence for the Buka area based on information from 73<br />

sites (Specht 1969: 211-213), spanning the period from Lapita to the present (Specht<br />

1969: Figure XII-52). Specht did not regard his sequence as continuous, being sure there<br />

were temporal sampling gaps. Specht used the hierarchical construct <strong>of</strong> tradition, style,<br />

substyle and attribute in his ceramic analysis (Gardin 1967, Willey & Phillips 1958),<br />

regarding attributes as irreducible units <strong>of</strong> description, with essential properties as<br />

discrete, mutually exclusive and immutable entities (Specht 1969: 64-65). Although he<br />

implicitly rejected the concept <strong>of</strong> pottery type as hindering the study <strong>of</strong> phyletics,<br />

regarding styles, in a materialist manner, by contrast, as involving time in their definition,<br />

his use <strong>of</strong> phyletic stylistic analysis was limited, as discussed further below, with styles<br />

being largely a finer-grained classification than types. Specht avoided multivariate<br />

computer classification approaches in favour <strong>of</strong> an iterative visual approach to analytical<br />

109


classification. He took a metric approach to vessel form variability, making use <strong>of</strong> vessel<br />

restriction factor, a measure <strong>of</strong> vessel form popular in American archaeology at that time.<br />

Specht included a succinct yet pertinent discussion <strong>of</strong> ceramic quantification, which<br />

anticipated Egl<strong>of</strong>f’s significant contribution in this area, suggesting that the subject had<br />

received some consideration at the ANU generally at that time, perhaps more so then than<br />

more recently (Egl<strong>of</strong>f 1973, Specht 1968:70-71, 2002). He used sherd count as his primary<br />

measure <strong>of</strong> quantity, but counted joining sherds as one. Where there were joins between<br />

layers, the combined sherd was counted in the layer with the most sherds/largest sherd,<br />

providing a rough correction for mixing in some instances.<br />

<strong>The</strong> recent end <strong>of</strong> the sequence was defined using surface-collected material. Here,<br />

Specht again provides a succinct and pertinent methodological discussion, this time <strong>of</strong><br />

seriation method, and suggests that unrestricted cliff-top site locations are more likely to<br />

provide short duration sites than those on the narrow coastal flat constrained by cliffs, the<br />

later being more likely to present mixed assemblages from extended or multiple periods <strong>of</strong><br />

occupation. He supported this theory with a number <strong>of</strong> stylistically homogenous collections<br />

from the cliff tops thought to be “single-period” occupations (Specht 1968:160-161).<br />

Specht placed heavy emphasis on paste classes in his analysis, an approach<br />

followed more recently on materials from the same general region by Summerhayes, and<br />

by Wickler and by Spriggs (Spriggs 1991, Summerhayes 1996, Wickler 2001) and in line<br />

with Shepard’s emphasis on technological attribute analyses (Shepard 1963).This approach<br />

has been partially borne out by Wickler’s subsequent analysis, in that the Lapita-age<br />

pottery largely contrasted with all other groups in this respect, being calcite-tempered,<br />

where all subsequent pottery had a significant terrigenous component, apart from<br />

occasional post-Lapita Sohano-style sherds (Specht 1968:192-194).<br />

Specht constructed a sequence <strong>of</strong> styles and sub-styles, beginning with slab-<br />

110


constructed, shell tempered Buka style, followed by Sohano one-piece paddle and anvil<br />

built style, (considered to have been strip-built, followed by paddle and anvil forming:<br />

Specht 1969:196). Hangan, Malasang, Mararing and Recent styles were considered a<br />

gradual evolution from Sohano, but Specht was equivocal on whether homologous<br />

similarity/heritable continuity and historical continuity could be seen between Buka<br />

(Lapita-age) and Sohano phases (Specht 1968 :195, 214, 216, 230). On the one hand some<br />

early Sohano substyle vessels occurred in crushed shell paste, and use <strong>of</strong> the paddle and<br />

anvil is a linking feature, but abandonment <strong>of</strong> slab construction, and the absence <strong>of</strong> a<br />

Watom-Lapita-style <strong>of</strong> broad rim notching from the large Sohano and later assemblages<br />

suggested a lack <strong>of</strong> heritable continuity.<br />

Some <strong>of</strong> Specht’s complex motifs (e.g. M34) diagnostic <strong>of</strong> particular substyles, are<br />

listed as occurring in other styles also, and this seems to indicate a too-direct reading <strong>of</strong><br />

stratigraphic superposition as recording changes in ceramic production. I think it more<br />

likely that complex motifs like M34 have a relatively specific period <strong>of</strong> production, and that<br />

these occur in other levels as a result <strong>of</strong> the vagaries <strong>of</strong> site formation. Just where this<br />

leaves Specht’s construction <strong>of</strong> a gradual evolution from Sohano to Recent is difficult to<br />

assess (see for example Specht 1969:258)<br />

Wickler on Buka:<br />

Wickler’s recent re-appraisal <strong>of</strong> the Buka/Nissan sequence, which largely reiterated<br />

Specht’s and Spriggs’ ceramic sequences with minor refinement, added substantial<br />

additional Lapita-phase data, resulting largely from inclusion <strong>of</strong> reef-flat locations in his<br />

survey strategy (Wickler 2001:8).<br />

Numerically large (high sherd count) yet quite broken assemblages <strong>of</strong> ceramics<br />

were recovered from Lapita reef sites and terrestrial test excavation <strong>The</strong> average sherd<br />

weight for a sample from site DAF was five grams, although sherds on the outer reef at<br />

111


this site were larger. At site DJQ 7.6% <strong>of</strong> sherds were over 6cm maximum dimension<br />

(Wickler 2001:76).<br />

Attribute description <strong>of</strong> Lapita sherds was made with the aims <strong>of</strong> investigating<br />

assemblage variability, and <strong>of</strong> seriating assemblages using a temporally sensitive subset <strong>of</strong><br />

attributes. Vessel forms were defined primarily on the basis <strong>of</strong> rim attributes (Wickler<br />

2001:77), but the criteria for distinguishing between everted rims <strong>of</strong> restricted vessels and<br />

open bowls are not clear. Wickler’s finding that the Lapita assemblage from site DJQ was<br />

dominated by open bowls (Wickler 2001:90) is difficult to evaluate in the absence <strong>of</strong><br />

illustrated examples <strong>of</strong> sherds that were used to make such identifications. Similarly,<br />

functional distinctions between assemblages based on vessel form (Wickler 2001:122) are<br />

not easily verified. No raw attribute or decorative data at the level <strong>of</strong> the sherd were<br />

presented, nor are examples <strong>of</strong> the sherds identifying some vessel forms illustrated. Some<br />

vessel form categories are very broad (“Form 9" for example). Sherd size was not recorded<br />

on a sherd-by-sherd basis, but for rim sherds an approximation <strong>of</strong> sherd size could be<br />

generated from the attributes “estimated rim diameter” and “rim percentage. <strong>The</strong> lack <strong>of</strong><br />

sherd size information makes identification <strong>of</strong> taphonomic processes and motif difficult.<br />

Lip form data was strikingly patterned by site, with the plain “Form 1" lip dominant<br />

at site DES (on Nissan) (commonly found on “Form 1a bowls” and “Form 4 bowls”).<br />

“Form 5" lip dominated at site DJQ (suggested to be diagnostic <strong>of</strong> “Form 2b” everted<br />

bowls), and “Form 11" lips were dominant in the large DAF sample (diagnostic <strong>of</strong> “Form<br />

9b” necked jars) (Wickler 2001:91). Continuous attribute interval class frequency data<br />

(Wickler 2001:92, Table 4.8) provides useful information on the ceramic sample size and<br />

vessel forms <strong>of</strong> the Lapita sites, with counts <strong>of</strong> neck angle classes, carination angle classes,<br />

and rim percentage classes, but it is impossible to know whether these counts are<br />

independent observations.<br />

112


Macroscopic temper analysis used a system <strong>of</strong> relative abundance ranking (Wickler<br />

2001:98) very similar to the final scheme adopted in Chapter 4 for the Roviana Data.<br />

Ferromagnesian tempers do occur in the Buka Lapita sites in low frequencies, so Specht’s<br />

finding that Lapita-age sherds were exclusively shell tempered was a sampling error. Seven<br />

sherds out <strong>of</strong> the 6519 for which temper class was identified had quartzo-feldspathic<br />

temper dominant, which may turn out to be more significant that initially thought by<br />

Wickler, as sherds <strong>of</strong> this temper type produced arresting petrographic results in the<br />

Roviana case (Dickinson 2000a, Felgate & Dickinson 2001). <strong>The</strong>se sherds were thought<br />

to be local by Wickler, but should be compared with the quartz-calcite hybrid tempers in<br />

the present study.<br />

His starting point in seriating the reef Lapita sites was to use frequency <strong>of</strong><br />

decoration, and frequency <strong>of</strong> dentate decoration, both <strong>of</strong> these characteristics thought to<br />

be indicators <strong>of</strong> early assemblages, based on the work <strong>of</strong> Green, Donovan, Parker, also <strong>of</strong><br />

the Arawe assemblages from the Bismarcks as outlined by Gosden and others, and also as<br />

in Anson’s analysis. Wickler’s seriation thus had an evolutionary component as the<br />

direction <strong>of</strong> change was assumed to be known. This may turn out to be a self-fulfilling<br />

prophesy, in light <strong>of</strong> Kirch’s recent suggestion <strong>of</strong> an early plain assemblage predating<br />

Lapita from Mussau.<br />

Wickler’s attribute frequency seriation <strong>of</strong> Lapita reef sites used an unusual and<br />

potentially problematic set <strong>of</strong> relative frequencies: frequency <strong>of</strong> decorated sherds<br />

(expressed as a percentage <strong>of</strong> the total sherd count), the frequency <strong>of</strong> dentate-stamping<br />

(again expressed as a percentage <strong>of</strong> total decoration?), bounded incision, and unbounded<br />

incision (these are also assumed to be percentages <strong>of</strong> the total number <strong>of</strong> decorated sherds<br />

as the rows in the seriation do not sum to 100). He found a low overall frequency <strong>of</strong><br />

decoration on the beach at DAF, with a high frequency <strong>of</strong> unbounded incised decoration,<br />

and characterized this a late site (300-100BC), while DJQ was regarded as falling within<br />

113


the Lapita period envisaged by Sand, 1000-800BC, with the highest frequency <strong>of</strong><br />

decoration, lots <strong>of</strong> dentate-stamping, and bounded rather than unbounded incision (Wickler<br />

2001:122). As discussed in the methodological reviews <strong>of</strong> quantification and seriation in<br />

Chapter 1, the percentage <strong>of</strong> sherds with decoration is easily biased by differences in<br />

brokenness (and brokenness is clearly different across his seriation units), while percentage<br />

decorated is also an attribute that may not seriate well in the first place due to the<br />

possibility <strong>of</strong> decorative extent and abundance over time.<br />

Wickler was careful to note the difficulty <strong>of</strong> separating functional and<br />

chronological variation, but felt that functional variation could only account in part for the<br />

reef site decorative/form variation. Also, the total number <strong>of</strong> collected samples was five<br />

(three areas from DAF and one each from DES and DJQ), which is on the light side for<br />

assessing the degree <strong>of</strong> temporal mixing. <strong>The</strong> possibility <strong>of</strong> temporal mixing with other<br />

later ceramic styles will be discussed further in Chapter 13.<br />

Wickler’s Lapita design analysis was a hybrid <strong>of</strong> Mead’s and Anson’s analytical<br />

schemes (Wickler 2001:122-127), with an additional feature <strong>of</strong> careful structure in relation<br />

to decorative location and details <strong>of</strong> execution (Wickler 2001: Appendix 1). By using<br />

Mead’s conception <strong>of</strong> design zone, and Mead’s concept <strong>of</strong> General/Restricted Zone<br />

Markers (Mead 1975: 26, Fig 2.11, 2.12), some conflation <strong>of</strong> the categories <strong>of</strong> decorative<br />

technique and decorative layout was introduced, creating proliferation <strong>of</strong> descriptive terms.<br />

<strong>The</strong> careful control for decorative location in his motif frequency analysis (Wickler 2001:<br />

Appendix A) is a welcome innovation to Lapita analysis (Mead’s design zones are<br />

locationally vague by comparison), but structural covariation <strong>of</strong> decoration <strong>of</strong> various<br />

vessel parts was not investigated in any detail or made use <strong>of</strong> in analysis. His motif<br />

occurrence and motif frequency comparisons were very much in the mode <strong>of</strong> Anson.<br />

Decorative classification did not take account <strong>of</strong> the structuring <strong>of</strong> decoration across the<br />

vessel.<br />

114


In reporting his most common Lapita stamped and incised motif frequencies as<br />

Anson motifs 435, 2, 187/188 and 421 (Anson 1986: 189-256, Wickler 2001: 124-125),<br />

the effect <strong>of</strong> sherd size on assignment <strong>of</strong> these very similar motifs is not discussed. In view<br />

<strong>of</strong> the small sherd sizes reported, how does one differentiate between motifs 435, 297, 190,<br />

188, 187 or 434 when these are all made up <strong>of</strong> parallel lines? Illustration <strong>of</strong> examples on<br />

which such identifications were based would have clarified these issues, as would<br />

presentation <strong>of</strong> raw analytical data, including sherd sizes (not recorded individually), vessel<br />

parts represented on the sherd, vessel form ascription and motif ascription. This point is<br />

not entirely critical for comparison with the Roviana materials, as, in his decorative<br />

attribute analysis <strong>of</strong> Lapita, Wickler made a useful distinction between bounded and<br />

unbounded incision, compatible with the present Roviana units <strong>of</strong> classification as<br />

discussed in Chapters 1, 4 and 9.<br />

Wickler’s Robinson Coefficient motif frequency comparison <strong>of</strong> Buka reef sites with<br />

Lapita sites <strong>of</strong> the Reef/Santa Cruz area, Watom and New Caledonia found the reef sites<br />

DJQ and DAF much more similar to each other than to any <strong>of</strong> the other sites, while DES<br />

(Tarmon on Nissan) was about as similar to RF2 and RF6 as to the other sites. <strong>The</strong> Jaccard<br />

coefficient motif sharing analysis using agreement scores gave different results, and<br />

Wickler was inclined to favour the frequency data (Robinson similarity coefficient) over<br />

the occurrence data. DES had a low motif count, so sample error was a possibility there<br />

(Wickler 2001:128-129). This general finding contrasts with Summerhayes’ parallel<br />

changes across the Lapita distribution from early Lapita to middle Lapita, unless the Buka<br />

Lapita is all very late, and thus more regionalized (in which case the presence <strong>of</strong> flat bases,<br />

stands, and some fine needle-like dentate conflicts with the Summerhayes/Anson<br />

hypothesis).<br />

Wickler’s analysis <strong>of</strong> post-Lapita excavated assemblages will not be reviewed, for<br />

reasons <strong>of</strong> brevity, since his findings did not differ from those <strong>of</strong> Specht. Wickler’s<br />

115


finding that there was evidence for heritable continuity between Lapita and Sohano phase<br />

is significant in that it links Lapita and post-Lapita in a tenuous way (as discussed by<br />

Specht, despite his eventual opting for no connection). Also Sohano-phase ceramics are<br />

the only post-Lapita Buka ceramics that bear any resemblance to the Roviana intertidal<br />

ceramics, although late-prehistoric Buka ceramics could be usefully compared to late<br />

Roviana terrestrial plainware (some <strong>of</strong> the Roviana “post-Lapita” styles may be present in<br />

the reef-flat collections <strong>of</strong> Wickler in low frequencies: this will be discussed further in<br />

Chapter 13).<br />

Regarding the nature <strong>of</strong> the transition from Buka (Lapita-Phase) to Sohano<br />

ceramics, he concluded:<br />

“Although handicapped by low sample size, disturbed deposits and a lack<br />

<strong>of</strong> reliable radiocarbon dates, the available evidence indicates a temporal<br />

overlap in the production <strong>of</strong> Buka (Lapita) and Sohano style ceramics, and<br />

a gradual replacement <strong>of</strong> the former by the latter (Wickler 2001 :144). ”<br />

While occasional compositional overlap and a few sherds <strong>of</strong> transitional style hinted at<br />

heritable continuity (Wickler 2001:141) the interpretation <strong>of</strong> this as gradual change, by<br />

default, in the absence <strong>of</strong> evidence to the contrary (Wickler 2001:168) cannot be regarded<br />

as evidence delimiting rates <strong>of</strong> ceramic change, and I take Wickler’s statements to mean<br />

that Buka and Sohano are related traditions, without necessarily accepting his suggestions<br />

<strong>of</strong> a gradual transition or temporal overlap in production, for which there seems to be no<br />

particular evidence in the detail <strong>of</strong> the excavations, where ample evidence for mixing <strong>of</strong><br />

ceramics across levels is presented.<br />

Vanuatu:<br />

This review will focus on the recent work by Bedford, in which earlier works are reviewed.<br />

Although a number <strong>of</strong> surface exposures <strong>of</strong> Lapita pottery have been recorded on Malo,<br />

116


Hedrick, who recorded many <strong>of</strong> these sites, focused on investigating settlement strategy<br />

through site location (Hedrick n.d:225-243), and, despite working with large surface<br />

samples, did not draw any detailed conclusions regarding temporal changes in Lapita<br />

decoration. Galipaud has conducted extensive surveys and limited excavation on Malo,<br />

Santo and Torres (Galipaud 1998), but has yet to advance any detailed ideas on the nature<br />

<strong>of</strong> temporal changes within the Lapita phase.<br />

Bedford in Vanuatu:<br />

A recent extensive reworking <strong>of</strong> the Vanuatu sequence by Bedford, which included<br />

reporting <strong>of</strong> substantial new excavation sequences, resolving a number <strong>of</strong> primary culture-<br />

historical issues, has Lapita as the foundation pottery style (Bedford 2000: 1, 35-37, 49-50,<br />

81, 153-165, 239, ) rather than Mangaasi (Bedford 2000:27). Subdivision or explication<br />

<strong>of</strong> Lapita-phase temporal variation was not attempted by Bedford, due to the paucity <strong>of</strong><br />

Lapita-phase ceramics recovered from reconnaissance testpitting in landscapes <strong>of</strong>ten deeply<br />

covered by tephra. Bedford considered that all Lapita samples were at the late end <strong>of</strong> a<br />

Lapita “Horizon”, with initial Lapita settlement circa 3000BP, after consideration <strong>of</strong> the<br />

radiocarbon evidence and the coarse dentate-stamping on the sherds recovered, which in<br />

the Anson/Summerhayes model indicate late date in a Lapita series. It seems likely that<br />

evidence for settlement dates similar to New Caledonia will turn up eventually.<br />

On Erromango and Efate, Bedford concluded that local sequences were a product<br />

<strong>of</strong> continuous ceramic evolution out <strong>of</strong> Lapita, while on Malakula, a post-Lapita Malua<br />

phase also bore a homologous similarity to Lapita. On Malakula, however, in contrast to<br />

Erromango and Efate, late Chachara bullet shaped pots were regarded as potentially non-<br />

homologous having “no apparent antecedents” among the poorly understood surface<br />

collected wares from the period between Malua and Chachara (Bedford 2000:147, 241).<br />

Accordingly, four regional Vanuatu phyletic series can be constructed from Bedford’s<br />

117


analysis, as follows:<br />

• Erromango: Dentate Lapita >>Ponamla Plain >> Early Ifo excurvate vessels with<br />

Fingernail impression >>Late Ifo incurving rims with fingernail impression and<br />

other decoration>>heavy notched plainware<br />

• Efate: Arapus Lapita-style plainware >> Early Erueti >> Late Erueti >>Early<br />

Mangaasi >> Late Mangaasi (Mangaasi possibly appearing on Malakula in surface<br />

collections also, although this is unconfirmed)<br />

• Malakula 1: Dentate Lapita >> Malua >> ? (small surface collected sample)<br />

• Malakula 2: Chachara >> Naamboi<br />

It is important, in reviewing Bedford’s data and constructing these four phyletic series,<br />

to question their historical completeness. While we know a lot more than we did<br />

previously about ceramic change in Vanuatu, do we know how well the regional ceramic<br />

changes within these four series are sampled? Bedford points out some blank spots in the<br />

Malakula record. Elsewhere, mixing <strong>of</strong> materials across strata may be a factor obscuring<br />

poorly-sampled periods. In Ponamla Area A, for example, age-depth correlation and<br />

“evidence for stylistic change....confirmed the stratigraphic integrity <strong>of</strong> the site” (Bedford<br />

2000:43), but stratigraphic unconformity (Bedford 2000: Figure 3.4) is best read as<br />

indicating that some excavation and possibly erosion <strong>of</strong> Layer 2 and 3a at least has<br />

occurred prior to deposition <strong>of</strong> Layer 1, incorporating materials originally deposited there<br />

into Layer 1, and that “stratigraphic integrity “ should not be read as absence <strong>of</strong> mixing,<br />

which Bedford confirms (Bedford 2002). Also, some <strong>of</strong> the dated Layer 1 material<br />

overlaps at two standard deviations with material from Layer 5, which is not to say that<br />

118


it is mixed to the point where stylistic change cannot be detected, but which should<br />

encourage a cautious approach to the question <strong>of</strong> whether gradual change with height in<br />

the column directly indicates gradual change in ceramic production style over time.<br />

Phyletic blank spaces in the series could be obscured by mixtures from different periods,<br />

especially when plainware phases exist, as on Erromango in the post-Lapita-phase<br />

(Ponamla plain). A degree <strong>of</strong> mixing might be hard to distinguish from the plain component<br />

<strong>of</strong> overlying Early Ifo, other than by vessel form (not possible in the Ponamla plain phase,<br />

in which vessel form seems little different to the succeeding early-Ifo vessel form). Having<br />

the two site sequences to compare has allowed a more complete sequence than would<br />

otherwise be the case, but is there yet more variety to be discovered in additional sites in<br />

the future? If so, how much more? Also, as mentioned above, it seems likely that Lapita<br />

face designs as found on New Caledonia and in the Southeast Solomons and Fiji will turn<br />

up eventually in most areas <strong>of</strong> Vanuatu.<br />

New Caledonia:<br />

<strong>The</strong> New Caledonian Lapita/Post-Lapita ceramic sequence has been elaborated recently<br />

by the work <strong>of</strong> Sand and colleagues, in which dentate-stamped pottery dates favour the<br />

range 1100BC to 800BC (Sand 1998:25-28) (Sand’s table <strong>of</strong> dates shows a broader<br />

absolute spread for Lapita contexts but he appears to have compressed this at an<br />

interpretive level). Sand sees these materials as allowing study <strong>of</strong> the decorative<br />

characteristics <strong>of</strong> the local Lapita motif inventory, with the aim <strong>of</strong> testing Kirch’s<br />

“Southern Lapita “ hypothesis. Sand sees a progressive replacement <strong>of</strong> Lapita ceramic<br />

ware by other incised or applied decorated ceramic traditions, and development from<br />

Lapita <strong>of</strong> the Podtanean paddle-impressed wares and Puen Mangaasi-like incised ware.<br />

Sand considered that some <strong>of</strong> the non-dentate pottery assemblages and non-dentate<br />

components <strong>of</strong> assemblages comprise a “Lapita-associated ceramic-series” within the<br />

119


Kone period (Sand 1999:143). He notes that these pots are found in spatial association<br />

with dentate-stamped Lapita pots, and in sites postdating Lapita. This category comprises<br />

triangular incised “non Lapita”, a range <strong>of</strong> shell-impressed styles, and paddle-impressed<br />

Podtanean pottery. While the presence <strong>of</strong> Podtanean in most Lapita sites, and in the<br />

earliest stratigraphic context at the Vatcha site, are taken by Galipaud to indicate<br />

contemporaneity with Lapita from first settlement, Sand suspects that this style <strong>of</strong> pottery<br />

in many cases represents later habitation than Lapita (Sand 2000:146). Galipaud sees the<br />

absence <strong>of</strong> Lapita in some Podtanean sites as a contemporaneous functional difference,<br />

with Podtanean being a utilitarian Lapita-age vessel type (Galipaud 1996:303). Sand has<br />

late/post Lapita impressed/incised wares in the north differentiating from incised Puen<br />

wares in the south, within his early Kone period (Sand 2000:155).<br />

Sand notes that ceramics excavated are still under analysis, but presents a<br />

preliminary synthesis <strong>of</strong> Kirch’s proposed “Southern Lapita Province”, defined by Kirch<br />

on the presence <strong>of</strong> the “Podtanean” paddle impressed wares in association with dentate-<br />

stamped Lapita ceramics (Kirch 1997:73). Sand proposes a short chronology for Lapita<br />

in this Southern province, beginning not earlier than 1100BC (Sand 1997a), and with<br />

dentate-stamping out <strong>of</strong> use by 750BC, a similar chronology to that recently proposed for<br />

the Eastern area <strong>of</strong> Fiji-western Polynesia. This is substantially shorter than the 1000 year<br />

chronology proposed for Lapita in the Bismarcks (Kirch 1997), although that chronology<br />

is in turn showing signs <strong>of</strong> contracting, with a figure <strong>of</strong> 500 to 600 years suggested more<br />

recently (Spriggs 2001).<br />

Sand saw “Southern Lapita” as having most direct links to the rest <strong>of</strong> the<br />

Melanesian chain, and having few links to less diverse eastern Lapita, which suggests that<br />

“southern” may be a misnomer for the more widespread characteristics, and “Central<br />

Lapita” might be a better term. Sand identified the major presence <strong>of</strong> composite rims on<br />

carinated pots as a Southern Lapita typological differentiation, but also noted a similar<br />

120


presence in Buka sites (Wickler 2001) (see comments above on Wickler’s identification<br />

<strong>of</strong> these as occurring on bowl vessel forms). <strong>The</strong>se composite rims will be shown to be a<br />

feature <strong>of</strong> one <strong>of</strong> the collection samples in the present study, so use <strong>of</strong> these as a defining<br />

characteristic <strong>of</strong> the Southern Lapita Series is significant to external comparison <strong>of</strong> the<br />

Roviana Lapita (see Chapter 13). Sand notes a varied set <strong>of</strong> tooth sizes in dentate-stamping<br />

from single sites, which contrasts with the Anson/Summerhayes “Early Far Western” fine-<br />

dentate hypothesis in the same way as Wickler’s Buka reef site data did. Sand notes that<br />

some parts <strong>of</strong> sites have almost all vessels decorated, while others have a relatively low<br />

percentage. He makes a distinction between geometric and anthropomorphic designs, and<br />

includes the stylized evolutions <strong>of</strong> anthropomorphic designs in the latter category. Of a<br />

range <strong>of</strong> complex geometric motifs, he describes one geometric motif as “characteristic <strong>of</strong><br />

the end <strong>of</strong> the Lapita period”. This motif is not illustrated, but is described as “the imprint<br />

<strong>of</strong> successions <strong>of</strong> straight lines, mostly done with a u-shaped tool”.<br />

Of the anthropomorphic motifs, he regards “double face motifs” as rare and found<br />

only in the earliest levels, and simpler than those found in the Western Lapita province. He<br />

notes that some face motifs (his “Type 3") are found in the Western and Far-Western<br />

provinces also, but not on cylinder stands and open pedestal bowls as in those provinces,<br />

rather on carinated pots, and occasionally plates with feet. In common with Spriggs<br />

(1990), Sand sees an evolution to an eye-nose-eye simplification, with final simplification<br />

being the loss <strong>of</strong> the eye, but does not support this with stratigraphic or other temporal<br />

evidence. In contrast, he states that this “evolved” motif is common from the earliest<br />

levels, indicating a foundation style or rapid local development, but regards these “stylized<br />

faces” as characteristic <strong>of</strong> the “Southern Lapita” dentate inventory. <strong>The</strong> rapid evolution<br />

from the “early” faces to their “later” derivatives in these New Caledonian sites puts the<br />

evolutionary seriations <strong>of</strong> Spriggs, Best and Ishimura under some pressure. Are the<br />

121


“earliest levels” in these sites representing substantial periods <strong>of</strong> time, <strong>of</strong> the order <strong>of</strong> a<br />

century or so, allowing this evolution to occur, or are the supposedly phyletic variants <strong>of</strong><br />

the faces contemporaneous? <strong>The</strong> way to rule out contemporaneity is to either find a<br />

sequence with the variants conveniently arranged in levels (Best does not seem to have<br />

evidence <strong>of</strong> this sort in his recent exposition)(Best 2002) or to seriate a good number <strong>of</strong><br />

large-sample low-diversify assemblages to create a fine-grained chronology, to test the<br />

chronology with some (preferably direct) dates, and to examine the relative positions <strong>of</strong><br />

the variants in the chronology. (Ishimura suggests that Frimiggacci has recorded stratified<br />

evidence for such change, but I have not had the opportunity to examine this).<br />

Non-dentate pots are also characteristic <strong>of</strong> the “earliest levels” (incised, shell-<br />

impressed, plain and paddle impressed). Incised decoration occurs on carinated pots, <strong>of</strong>ten<br />

with a notched rim. Rounded triangles in a frieze are common. <strong>The</strong>se motifs drop out “at<br />

the same time as the dentate-stamped motifs”, and do not occur later, supporting in my<br />

opinion a view <strong>of</strong> Lapita as a stylistic horizon rather than series on current published<br />

evidence. This opinion is given tentatively, in the absence <strong>of</strong> published detailed ceramic<br />

analyses from the recent New Caledonian excavations, but it should be noted that this is<br />

a major contrast to the series proposed for the Bismarcks by Summerhayes. Incised motifs<br />

are said to be strikingly less diverse than those in Western/Far Western Lapita. Shell<br />

impressions are uncommon but are present in all major Lapita sites. Notched lips or<br />

entirely plain pots are mostly carinated with excurvate rim, but some are simpler<br />

uncarinated forms.<br />

Paddle impressions are restricted to carinated forms with outcurved rims (although<br />

Sand’s “carination” as illustrated is a substantially wider-angle affair than that illustrated<br />

for dentate carinations). Paddle-impressed decoration is said to occur on better quality<br />

paste with thinner and harder pots. Paddle-impression occurs in low percentage in “early”<br />

sites; “...their major development takes place after the demise <strong>of</strong> the dentate-stamped and<br />

122


incised Lapita motifs....” Sand does not explicitly discuss the role <strong>of</strong> mixing and issues <strong>of</strong><br />

occupation span in the association between Podtanean and early dentate Lapita but, he<br />

gives examples where the dentate and Podtanean are separate and superposed, e.g. in<br />

rockshelter LWT008 <strong>of</strong> Hnajoisisi <strong>of</strong> Lifou island, while elsewhere intermediate forms are<br />

found. <strong>The</strong>re seems then to be still an open question <strong>of</strong> whether Podtenaean evolved<br />

gradually out <strong>of</strong> Lapita or whether the transition, <strong>of</strong> whatever kind, was relatively sudden,<br />

and blurred in general by mixed deposition and lengthy occupation spans.<br />

<strong>The</strong> View from the East: Fiji and Tonga:<br />

Best constructed an attribute-based Robinson coefficient similarity matrix seriation for the<br />

entire Fijian sequence based on evidence from stratified excavations on Lakeba,<br />

supplemented by other excavated and surface collected assemblages. For the Lapita period<br />

his data comprised the Natunuku and Naigani excavations, and the stratified Lakeba Site<br />

196 (an extensive and rich open site thought to have a stratigraphic sequence <strong>of</strong> extended<br />

duration with some temporal gaps and some mixing). Within the Lapita “style”,<br />

“...the number <strong>of</strong> vessel shapes reduce from 12 to six, carinated forms<br />

vanish, and decoration declines from initial fairly complex designs to simple<br />

arcs and zig-zags along the rims (Best 2002:17).”<br />

<strong>The</strong> chronological trend within the Lapita period noted by Best, the loss <strong>of</strong> carinated<br />

vessel forms, is not in evidence in Best’s data from site 196, where these occur in all levels<br />

<strong>of</strong> the area excavation squares (Best 1984:ppA3-A4). Strictly speaking, some carinated<br />

forms persist in the later levels <strong>of</strong> Lakeba site 196, and it is dentate-decorated sharp<br />

carinations which drop out rapidly from the lowest levels (Best 1984: table <strong>of</strong> attribute<br />

combinations for site 196).<br />

While the Lakeba excavations provide convincing evidence for a shift to arc-<br />

impressed or arc-stamped rims over time, Best’s data on Lapita motifs is not as clear-cut.<br />

A complicating factor making the argument difficult to abstract from the data is the use<br />

123


<strong>of</strong> two separate systems <strong>of</strong> motif description/classification: the Mead system and the<br />

Frost-Irwin system. Arc rims are coded in the Frost-Irwin manner as attribute<br />

combinations, while the Mead-system motif classes are omitted from the seriations.<br />

<strong>The</strong> Mead-system data from excavations at Lakeba site 196 is quantified by sherd<br />

count in an appendix, and does not show clear trends by level, beyond what might be<br />

expected from the motif sample sizes (Best 1984:A21-A22). Best stresses that if the counts<br />

<strong>of</strong> arc rims and plain sherds were added into these data a clearer trend would emerge<br />

(Best, personal communication 2002), but there is no clear motif shift by level using the<br />

Mead motif data alone.<br />

In his recent synthesis which has roulette stamping and “face” motifs early in the<br />

Lapita series, simpler Mead-type motifs incorporating design elements such as DE1 and<br />

DE2 (Best 1984) persisting from early Eastern Lapita to some sort <strong>of</strong> mid-Lapita, and arc-<br />

rims as the last stage <strong>of</strong> Lapita, the supporting data for the first two stages <strong>of</strong> this transition<br />

are seen in the similarity on the one hand <strong>of</strong> Natunuku and Naigani and the Reef/Santa-<br />

Cruz sites (with high frequency <strong>of</strong> “face” sherds and Mead Motif M.33), and on the other<br />

hand the Lakeba open site and most Tongan Lapita sites, being “late” with the complex<br />

anthropomorphic motifs not found or else explicable as heirloom effect.<br />

<strong>The</strong> low frequency <strong>of</strong> “early” motifs at Lakeba site 196 may be a sampling error.<br />

Although 2000 sherds were recovered from the site, from a scatter <strong>of</strong> 17 test pits and two<br />

groups <strong>of</strong> excavation squares each totaling 12m 2 (Best 1984:figure 2.16), the 15000m 2<br />

open site may hide considerable spatial diversity in the large unexcavated areas.<br />

Best’s suggestion in “a View from the East” (2002) that motif M12.2 is common<br />

at Lakeba site 196 and can be traced through a sequence <strong>of</strong> decreasing complexity if<br />

examples are contrasted with sherds from the Reef Islands is a phyletic or evolutionary<br />

seriation that is not supported stratigraphically, as the sherd from square 12 layer A1 <strong>of</strong><br />

site 196 has seven underlying occurrences <strong>of</strong> the same motif, mostly from the basal layers<br />

124


C and B3, none <strong>of</strong> which are illustrated in his “series”. Surely these sherds are relevant to<br />

the proposed series? Picking a phyletic series from the vast area which includes the Reef<br />

Islands and Fiji loses control <strong>of</strong> sources <strong>of</strong> variability, and becomes a re-statement <strong>of</strong> the<br />

definition <strong>of</strong> Lapita, that Lapita simplifies over time and becomes non-Lapita, rather than<br />

being a demonstration or confirmation <strong>of</strong> the way that change happened.<br />

While Best demonstrates convincingly that a specific carination form is present only<br />

in the lowest levels <strong>of</strong> Lakeba, and Best’s recent evolutionary simplification seriation may<br />

well be correct for Fiji, and even for the entire Lapita distribution, like Spriggs'<br />

formulation, it suffers from a circularity <strong>of</strong> reasoning. <strong>The</strong> direction <strong>of</strong> change is known<br />

and is the same everywhere for Lapita, therefore we can pick sherds out <strong>of</strong> sites and order<br />

them in a series, therefore we can order sites in time, and by doing this we can show that<br />

changes occur in concert across the Lapita distribution, therefore interaction is happening,<br />

as similar changes occur everywhere.<br />

Best’s proposal that particular Mead motifs can be used to date Lapita sites might,<br />

like Summerhayes’ model, be attributing the utilitarian component <strong>of</strong> sites to a late date,<br />

and the fancy component to early Lapita. Rounded carinations are present throughout the<br />

Lakeba strata, yet an excavation which finds the rounded carination form but not the sharp<br />

carination form with dentate is liable to be regarded as late (I am not referring here to the<br />

arc-rim post-Lapita decoration, which clearly supersedes Lapita at site 196).<br />

Lapita Temporal Variability -Conclusions:<br />

For West New Britain and Anir I have argued in detail above that Summerhayes’ chrono-<br />

stratigraphic decorative trends are problematic, and that higher-level inferences regarding<br />

Lapita interaction and a universal Lapita series need to accommodate these uncertainties.<br />

This critique was made possible by Summerhayes’ thorough sherds-as-vessels analysis,<br />

125


combined with Gosden and Webb’s geomorphological data. <strong>The</strong>se trends should not be<br />

regarded as confirming a regional West-New Britain gradual transition from dentate<br />

stamped decoration to incised, but should rather continue to have the status <strong>of</strong> hypothesis,<br />

in that they tell us little more than that at first there was Lapita and then there was<br />

something else. A correlation between vessel form and dentate-stamping has been<br />

demonstrated, but whether this varies temporally is less secure.<br />

Summerhayes’ motif-occurrence/motif sharing clustering/PCA comparisons with<br />

the wider Lapita world were dogged by a fundamental mismatch between the systematics<br />

<strong>of</strong> Anson's splitting classificatory approach (which results in sparse data <strong>of</strong> unknown<br />

representativeness), and the motif occurrence grouping methods used. Even if the<br />

dichotomy in the data is real, is the explanation temporal? Summerhayes’ recent<br />

chronological hygiene exercise for the Arawe and Anir dates is not conducted independent<br />

<strong>of</strong> the ceramic temporal hypothesis, and is not independent confirmation <strong>of</strong> the<br />

Early/Middle Lapita chronology in its present form, and is better regarded as a preferred<br />

explanation. Some late dates from “early” contexts were rejected because they did not fit<br />

the percieved pattern <strong>of</strong> temporal ceramic variability..<br />

Watom: the potentially broad age-span <strong>of</strong> materials from the upper C1 SAC<br />

excavation zone means that the recovered ceramics from zone CI and C2 are not <strong>of</strong> an<br />

accurately known age, Also, the spread <strong>of</strong> ages can be expected to differ by level at SAC,<br />

to an extent that suggests Watom does not yet provide a secure well-tested construct <strong>of</strong><br />

Lapita temporal variability against which Roviana variability can be compared.<br />

Mussau: Detailed results <strong>of</strong> ceramic analyses are as yet unpublished, and Kirch’s<br />

claim for stratigraphic evidence for a progression from plain to dentate to incised<br />

decoration is subject to similar C14 and formation-process related criticisms to those<br />

directed at Summerhayes’ Arawe analyses. Formation-based criticisms are principally that<br />

sub-tidal mixing <strong>of</strong> area B zone C may have occurred to some extent prior to burial, and<br />

126


that superposed incised pottery in Zone B may ultimately be derived from functionally or<br />

taphonomically differentiated areas <strong>of</strong> the sub-tidal deposit, cast up on the beach at a later<br />

date. Until detailed information on vessel joins, appropriateness <strong>of</strong> units <strong>of</strong> quantification<br />

etc is available, preliminary ceramic conclusions from Mussau do not provide a detailed<br />

construct <strong>of</strong> temporal variability to compare the Roviana ceramics with.<br />

Admiralties: Wahome’s seriation <strong>of</strong> a large number <strong>of</strong> Admiralties sites, including<br />

excavated and surface-collected units, provides a four-period chronology that on the<br />

surface avoids many <strong>of</strong> the pitfalls <strong>of</strong> seriation by attention to the statistical basics. <strong>The</strong> 700<br />

year gap suggested by Ambrose limits the utility <strong>of</strong> Wahome’s sequence for comparisons<br />

in the present context.<br />

Buka: Specht’s chronology from Buka to later styles, amplified by Wickler to<br />

include more Lapita variability as evidenced by three reef-flat sites (one from Nissan), and<br />

to tenuously link Lapita to Sohano, provided the nearest and most comprehensive ceramic<br />

chronology <strong>of</strong> relevance both temporally, geographically, and in terms <strong>of</strong> formation<br />

process, to the Roviana materials. His seriation <strong>of</strong> a small number <strong>of</strong> ceramic samples<br />

encompasses, in his view, functional variability between assemblages indicated by vessel<br />

form, and also possibly taphonomic variability in that three <strong>of</strong> his assemblages, from<br />

various parts <strong>of</strong> site DAF, represent taphonomic variability between reef edge, reef flat,<br />

and coastal-erosion scatter, the latter from what was regarded as formerly a terrestrial<br />

deposit. DES from Nissan falls into this latter category too, according to Spriggs (1991).<br />

Wickler’s seriation also faces the “number <strong>of</strong> sites” sample size problem (as does the<br />

current Roviana data) and suffers from the use <strong>of</strong> a ratio <strong>of</strong> plain:decorated, a measure<br />

easily biased by differences in brokenness.<br />

Wickler’s motif frequency grouping <strong>of</strong> Buka and other assemblages does not<br />

accord well with Summerhayes’ universal Lapita ceramic series. Despite Wickler’s<br />

hypothesis that the sites represented some time depth, those with reasonable sample sizes<br />

127


were much more similar to each other than to either Watom, the RSZ sites or Ile des Pins,<br />

suggesting a greater degree <strong>of</strong> regionalization than predicted by the Summerhayes model.<br />

Reef / Santa Cruz: use <strong>of</strong> Jaccard coefficient on sparse data <strong>of</strong> unknown<br />

representativeness/sample size (the latter resulting from choice <strong>of</strong> units <strong>of</strong> quantification)<br />

is a flaw in Green’s (1978) seriations. Spriggs, in his temporal hypothesis <strong>of</strong> the changing<br />

face <strong>of</strong> Lapita, suggested that a variety <strong>of</strong> face design types were present at both SZ-8 and<br />

RF-2, as did Ishimura, and this corresponds with my own reading <strong>of</strong> Donovan’s and<br />

Parker’s analyses, in which, in addition to the varied face designs, there seem to be<br />

incised/notched and other geometric linear Lapita variants not discussed by Spriggs, but<br />

potentially <strong>of</strong> chronological significance. Parker understood these sites as sequential slices<br />

in time, based primarily on the then radiocarbon evidence, as did Green in 1978, since<br />

revised (Green 1991c) and currently the subject <strong>of</strong> further dating research by Green and<br />

others. I have suggested that an alternative interpretation <strong>of</strong> these sites as roughly<br />

contemporaneous with variable, potentially overlapping occupation spans is consistent with<br />

the evidence. While C14 dates obtained so far from RF6 are younger than the other sites,<br />

I would think that at least some <strong>of</strong> the material in RF6 is <strong>of</strong> a similar age to SZ8/RF2,<br />

unless it can be demonstrated by C14 that such an overlap in occupation span is unlikely<br />

(the portion <strong>of</strong> the RF2 sample from the northern half <strong>of</strong> that site may be key here). Best’s<br />

suggestion that the site order should be reversed (Best 2002:93) is in my view not<br />

supported by current evidence, and also involves a sequential slice-in-time view <strong>of</strong> the<br />

sites.<br />

<strong>The</strong> evidence from Vanuatu suggests that rather than a Lapita ceramic series,<br />

evidence for ceramic variability does not presently extend beyond the general features <strong>of</strong><br />

the transition from Lapita to post-Lapita, and in this latter period diverse regional<br />

sequences are seen. <strong>The</strong> samples <strong>of</strong> Lapita recovered by recent testpitting reconnaissance<br />

and excavation are too small to allow a fine-grained temporal construction within<br />

128


Bedford’s Lapita phase. It is tempting to see parallels between some Erromango ceramics<br />

and some <strong>of</strong> the Roviana materials but these may be unrelated co-incidence, as will be<br />

discussed further in a Chapter 13 (see also Felgate 2002).<br />

New Caledonia is emerging as a region in which a process <strong>of</strong> Lapita-phase<br />

sequence building is made possible by a large sample <strong>of</strong> Lapita sites, and the only proviso<br />

here is that the descriptive records published to date lack the detail <strong>of</strong>, for example,<br />

Summerhayes’ Arawe sherd analyses. Comparisons can be usefully drawn between New<br />

Caledonian Lapita and Roviana ceramics from Honiavasa, as predicted by Sand’s<br />

comments regarding similarities <strong>of</strong> New Caledonian Lapita to Buka Lapita.<br />

This review <strong>of</strong> the Lapita ceramic series concludes that the regional Lapita ceramic<br />

series from New Britain, Mussau, Buka, and the Reef-Santa-Cruz sites should be<br />

characterized as untested hypotheses rather than securely established chronological<br />

frameworks, and that Lapita still constitutes a ceramic stylistic horizon rather than a series<br />

in the Western/Far Western region at least, for practical purposes <strong>of</strong> external comparisons<br />

<strong>of</strong> the Roviana materials. Publication <strong>of</strong> details <strong>of</strong> the Mussau ceramics would probably<br />

revise this assessment. <strong>The</strong> best samples for external comparisons are the Buka reef-site<br />

samples, since natural formation processes are so similar (wave processes are dominant).<br />

<strong>The</strong> terrestrial samples and buried stilt house samples are problematic in this respect, but<br />

also suffer from chronological uncertainties which make comparisons difficult. Aspects <strong>of</strong><br />

the Roviana ceramics bring to mind characteristics <strong>of</strong> Southern Lapita, and characteristics<br />

<strong>of</strong> Summerhayes’ early Lapita, but decoration is “Late” if Roviana Lapita is slotted into<br />

the Summerhayes model. <strong>The</strong>se issues will be covered in some detail in making external<br />

comparisons in Chapter 13; it is simply noted at this stage that some difficulties arise in<br />

trying to slot the Roviana pottery into any existing “Lapita ceramic series”. <strong>The</strong>se<br />

difficulties may arise from errors in the current formulations <strong>of</strong> those series rather than in<br />

idiosyncratic variability in the Roviana case.<br />

<strong>The</strong> impression given by the authors <strong>of</strong> many <strong>of</strong> these regional sequences (with the<br />

129


exception <strong>of</strong> Anson (1983) and Wickler (2001) is that these are all confirmation <strong>of</strong> an<br />

expected Lapita trend from dentate to incised/applied, and <strong>of</strong> the temporal primacy <strong>of</strong> the<br />

“Far Western” sites (Kirch mentions early plainware at Mussau recently though). <strong>The</strong>re is<br />

a circularity in this reasoning stemming from a tendency to recognize evidence for the null<br />

hypothesis, that Lapita became non-Lapita, as evidence for the details <strong>of</strong> the transition.<br />

I conclude specifically that some attributes <strong>of</strong> pottery samples are particularly likely<br />

to mislead in this regard, for example, using ratio <strong>of</strong> plain to decorated pottery, quantified<br />

by sherd count, in seriations. Smash the pottery up and it gets younger. More robust<br />

properties <strong>of</strong> ceramic samples, such as the number <strong>of</strong> shared motifs, are sensitive to<br />

sample-size differences, particularly when units <strong>of</strong> motif classification result in sparse data,<br />

requiring careful attention to sample evaluation, selection <strong>of</strong> units <strong>of</strong> quantification and<br />

method <strong>of</strong> analysis.<br />

This assessment <strong>of</strong> the nature <strong>of</strong> our current regional samples and the stability <strong>of</strong><br />

existing temporal constructs is not an argument against stratigraphic excavation, nor is it<br />

a condemnation <strong>of</strong> the value <strong>of</strong> ceramic chronology in general, or an attempt to denigrate<br />

the work <strong>of</strong> others. It is an assessment principally <strong>of</strong> the level <strong>of</strong> temporal resolution <strong>of</strong><br />

secure regional chronologies. I argue that while outlines <strong>of</strong> a culture-history are stable now<br />

for several island groups, we are running up against the low temporal resolution <strong>of</strong><br />

reconnaissance sampling and C14 dating-by-association. Research design <strong>of</strong> the Lapita<br />

Homeland Project focused on the identification and excavation <strong>of</strong> stratified sequences, as<br />

has recent work on Vanuatu. <strong>The</strong>re is nothing wrong with that, as long as it is accepted<br />

that it is difficult to assess within-site spatial structure, particularly for the Lapita period.<br />

<strong>The</strong>re is a dearth <strong>of</strong> large, representative samples <strong>of</strong> Lapita from Near Oceania,<br />

which means seriation studies, which are key to increasing the temporal resolution <strong>of</strong><br />

chronologies, are not succeeding as well as they might were better samples available. <strong>The</strong><br />

difficulty <strong>of</strong> obtaining fine-resolution stratified sequences may well be compounded by<br />

Lapita and post-Lapita settlement pattern instability, and low-deposition life ways<br />

130


compared to the archaeology <strong>of</strong> the Near East or American Southwest (post and thatch<br />

in an environment <strong>of</strong> rapid organic weathering rather than mud brick houses in a dry<br />

climate) leading to a dearth <strong>of</strong> well-stratified records <strong>of</strong> change unless there is a regular<br />

natural airfall tephra supply, as in much <strong>of</strong> Vanuatu.<br />

Our archaeological methods have yet to adapt to low-deposition conditions by<br />

focussing on the horizontal component <strong>of</strong> temporal patterning, rather than expecting a<br />

temporal layercake in settlements. Beyond deconstructing Summerhayes’s current<br />

universal Lapita Ceramic Series, this review identifies a need to search out and analyse<br />

archaeological distributions in horizontal space to a greater extent than has been the norm<br />

for Lapita in Near Oceania to date. <strong>The</strong> recovery and seriation <strong>of</strong> a large number <strong>of</strong> good<br />

short-term samples is key to boosting ceramic time-resolution. C14 dating, in this view,<br />

can be used as a confirmatory technique rather than the primary means by which ceramic<br />

chronology is constructed, especially with the rise <strong>of</strong> direct-dating using AMS.<br />

131


132


CHAPTER 3:<br />

SCALE AND METHOD OF FIELD SURVEY<br />

REQUIRED TO GENERATE A SAMPLE OF<br />

LAPITA SITES FOR SERIATION IN THE NEW<br />

Introduction:<br />

GEORGIA REGION<br />

It is traditional to preface the analytical portions <strong>of</strong> PhD theses with a <strong>chapter</strong> on the sites<br />

and their setting. Here, I try rather to take a problem-oriented analytical approach to the<br />

archaeological landscape <strong>of</strong> Near Oceania, in the course <strong>of</strong> which I hope I make clear the<br />

location and nature <strong>of</strong> surface collections used in other analytical <strong>chapter</strong>s. Detailed spatial<br />

analyses <strong>of</strong> sites are provided in Chapter 11.<br />

While the claim could be made that the Roviana intertidal survey data were the<br />

result <strong>of</strong> careful and systematic research design from the outset <strong>of</strong> the project, the prosaic<br />

version <strong>of</strong> events is that research design occurred as a series <strong>of</strong> ideas arising before and<br />

during the fieldwork. This is particularly true <strong>of</strong> the survey design, which began life as a<br />

search for pottery sites primarily through local knowledge (hereafter referred to a the<br />

informant-prospection method), and evolved to a more systematic inspection <strong>of</strong> the<br />

intertidal and subtidal zones once it became apparent that these were the locus <strong>of</strong> early<br />

ceramic evidence. Initial aims were simply site discovery. A concern with site visibility,<br />

survey coverage, site preservation and distributional patterning crystallised in the later<br />

stages <strong>of</strong> fieldwork, during the 1997-8 field seasons, when survey was focused in the<br />

Kaliquongu region, with the aim <strong>of</strong> characterization <strong>of</strong> the archaeological distribution in<br />

the intertidal zone. <strong>The</strong>se two distinct stages or types <strong>of</strong> survey will be referred to as the<br />

133


Roviana survey and the Kaliquongu survey respectively. During writing up, I saw a need<br />

to phrase the information in terms <strong>of</strong> an estimation approach and hypothesis testing, largely<br />

in order to demonstrate the limitations <strong>of</strong> the data in this regard, and to generate<br />

conclusions for future practice.<br />

While the following <strong>chapter</strong> might well be characterized as more and more about<br />

less and less, which in terms <strong>of</strong> the size <strong>of</strong> the region it is, this is not a bad thing; the results<br />

<strong>of</strong> iterative survey at an increasing level <strong>of</strong> detail are informative, and raise fundamental<br />

epistemological issues for the early ceramic period in Near-Oceania. <strong>The</strong> foremost <strong>of</strong> these<br />

concerns sampling at the level <strong>of</strong> regional survey. <strong>The</strong> Roviana Lagoon surveys are<br />

presented below within a theoretical framework <strong>of</strong> sample-surveying. <strong>The</strong> objectives <strong>of</strong> the<br />

surveys are first to test the hypothesis that early-Lapita was present in the New Georgia<br />

raised coral barrier-island/lagoon system in the past, second, to estimate site density for<br />

Lapita and post-Lapita ceramic-lithic intertidal scatters (using a ceramic chronology which<br />

will be justified in detail in succeeding <strong>chapter</strong>s), and third, to draw methodological<br />

conclusions about the scale <strong>of</strong> survey needed to obtain a sufficient site sample for seriation<br />

<strong>of</strong> Lapita sites. <strong>The</strong> characteristic <strong>of</strong> the local archaeological record that allows this<br />

approach is good site visibility on coralline lagoon shorelines in the intertidal zone. <strong>The</strong><br />

level <strong>of</strong> preservation evident in samples, implications for survey and analysis <strong>of</strong> these<br />

results, and what the collection sites represent in terms <strong>of</strong> settlements, etc are the subjects<br />

<strong>of</strong> the following three <strong>chapter</strong>s.<br />

Review <strong>of</strong> Landscape/settlement Pattern Studies in Oceania:<br />

Through the 1970s to 1990s a number <strong>of</strong> settlement-pattern studies were published, which<br />

involved coarse ceramic periodization and an implicit commitment to synonymy <strong>of</strong> “site”<br />

and “settlement”(Anderson 2002, Best 1984, Burley 1994, Hedrick n.d, Irwin 1972, 1985,<br />

Kirch 1988a, Lep<strong>of</strong>sky 1988, Sand 1997b, Swadling 1976). More recently, landscape<br />

approaches have been advocated, in which<br />

“structures <strong>of</strong> reference are made manifest not only through sites, but<br />

134


through broadly distributed scatters <strong>of</strong> material <strong>of</strong> the sort picked up by socalled<br />

“<strong>of</strong>f-site” approaches....understanding the overall distribution <strong>of</strong><br />

materials is dependent in turn on understanding the changes to which the<br />

landscape has been subject (Gosden & Head 1994)”.<br />

Approaches <strong>of</strong> this sort to surface scatters, in which formation processes, archaeological<br />

visibility and site preservation are controlled for have not been applied to the archaeology<br />

<strong>of</strong> Lapita pottery. This aspatial tendency <strong>of</strong> Lapita archaeological method, where the site,<br />

the layer assemblage, or even the test pit tends to be the primary unit <strong>of</strong> analysis, at the<br />

expense <strong>of</strong> regional distributional analysis (see Clark 1999:455 for a recent example)<br />

requires further discussion, as the Kaliquongu survey described below stands in contrast<br />

to this tendency. In near-Oceanic Lapita studies, much effort has been directed at finding<br />

what has been described in another context as,<br />

“<strong>The</strong> optimal solution to a culture-historical puzzle...a single point where<br />

all <strong>of</strong> a region’s typological complexes could be found on top <strong>of</strong> each<br />

other, separated by sterile deposits.” (Wobst 1983:42)<br />

In archaeology elsewhere prior to 1960 this emphasis worked together with a large-site<br />

bias to divert attention from issues <strong>of</strong> sampling and estimation, and limited archaeological<br />

survey to a prospecting method (Wobst 1983), and this characterization would seem to<br />

hold in near-Oceania also, particularly for the Lapita Homeland Project, but also for some<br />

more recent studies (Bedford 2000, Clark 1999).<br />

Wobst noted that seriation studies, in contrast to culture-historical excavation<br />

approaches, led to a search for short-term occupation sites within short distances <strong>of</strong> each<br />

other, which necessarily increased concern for intrasite distributions. Examples <strong>of</strong> this<br />

sort <strong>of</strong> approach to survey can be found in the works <strong>of</strong> Specht and also Irwin (Irwin 1972,<br />

1985, Specht 1969). Wobst criticised culture historical approaches as lacking an explicit<br />

rationale for excavation and survey, and for an inability to evaluate the accuracy,<br />

precision, and consistency <strong>of</strong> their results, and Gosden expressed similar sentiments in<br />

relation to Lapita in the Bismarck archipelago (Gosden 1991a). <strong>The</strong> recent use <strong>of</strong> site<br />

density data by Anderson in the course <strong>of</strong> assessing Lapita mobility suffers a similar fatal<br />

135


flaw, unmitigated by the argument that,<br />

“...the broad pattern <strong>of</strong> current distributional data has remained much the<br />

same for more than 20 years, suggesting that it is not just on the grounds<br />

<strong>of</strong> relative site recording intensity that the density <strong>of</strong> sites on the coasts <strong>of</strong><br />

large islands is lower than on small islands throughout the Lapita range, or<br />

that mainland New Guinea, and perhaps the main Solomon Islands, are<br />

largely bereft <strong>of</strong> Lapita sites (especially early Lapita sites) (Anderson<br />

2002).”<br />

Evidence <strong>of</strong> absence does not accumulate by default, nor does absence <strong>of</strong> evidence<br />

transform into evidence <strong>of</strong> absence with the passage <strong>of</strong> time (see Felgate 2002 for a<br />

discussion <strong>of</strong> the epistemology <strong>of</strong> the Lapita distribution in Near Oceania). It remains to<br />

be demonstrated that the recorded density <strong>of</strong> Lapita sites is a preserved settlement pattern.<br />

I think this is highly unlikely, and showing why this is so is a major component <strong>of</strong> this<br />

thesis.<br />

Early approaches to archaeology as sampling have been described as “...something<br />

<strong>of</strong> a programmatic reaction to culture history rather than well justified applications to<br />

resolve important research questions.”(Wobst 1983:43) Between 1960 and 1975 Binford’s<br />

research design/sampling approach was applied predominantly in regions where there was<br />

excellent archaeological visibility on the surface (Wobst 1983:47). A feature <strong>of</strong> most such<br />

studies was stratified sampling using existing knowledge, and multi-stage survey strategies,<br />

although objectives seldom extended to specifying the wider ramifications for the<br />

distribution beyond the sample region.(Wobst 1983: 47-48). <strong>The</strong> introduction <strong>of</strong><br />

hypothesis-testing survey objectives in the mid 70s (for example Cowgill 1975:261)<br />

marked a further advance in sampling research design, but such developments were still<br />

confined to regions with good archaeological visibility. Wobst notes a trend since 1976 to<br />

extend the new methods to areas with poor surface visibility and a consequent emphasis<br />

on subsurface prospection techniques for such regions.<br />

Seen in the context <strong>of</strong> Wobst’s analysis <strong>of</strong> the development <strong>of</strong> sampling<br />

approaches to regional archaeology, a likely explanation for aspatial approaches to the<br />

study <strong>of</strong> the regional distribution <strong>of</strong> Lapita is the environmental context <strong>of</strong> the Lapita<br />

136


Homeland Project, the major contributor to the Lapita knowledge-base in near-Oceania.<br />

Holocene rhyolitic volcanism and tectonic instability in the Bismarck archipelago has<br />

created a landscape in which few archaeologists would have the temerity to claim the<br />

pristine preservation and visibility <strong>of</strong> Lapita social landscapes. Lapita coastal site<br />

distributions in this dynamic landscape are expected to be subject to a high degree <strong>of</strong><br />

obliteration or burial. Anderson points out that most Lapita sites are coastal, without any<br />

significant discussion <strong>of</strong> the impact <strong>of</strong> coastal geomorphological processes on site<br />

preservation and visibility. It cannot be merely<br />

“...accepted either that the archaeological patterning is approximately<br />

representative, or that proposition may be taken simply as a working<br />

hypothesis (Anderson 2002:16).”<br />

If the record is simply accepted as representative <strong>of</strong> a phenomenon as a working<br />

hypothesis, there can be no assessment <strong>of</strong> the confidence we can have in higher-level<br />

hypotheses built on that data. Resort to a recursive array <strong>of</strong> contingent untested hypotheses<br />

renders archaeology a continual tabula rasa for the idiosyncratic rewriting <strong>of</strong> higher level<br />

theory.<br />

<strong>The</strong> sampling sophistication necessitated by areas with poor preservation and/or<br />

poor surface visibility has been slow to develop in Lapita archaeology, and while it has<br />

been theorized to some extent (Gosden 1991b), the requisite sampling sophistication that<br />

would allow distributional analyses <strong>of</strong> Lapita is largely absent from existing treatments <strong>of</strong><br />

the data and approaches to the archaeology <strong>of</strong> the region and period. Thus the Binfordian<br />

exhortation that archaeology should promote the region to the position <strong>of</strong> the primary unit<br />

<strong>of</strong> archaeological research (Binford 1964) with explicit research design and explicit use <strong>of</strong><br />

sampling theory, has largely fallen on stony (tephra covered?) ground in near-Oceania,<br />

where the site has remained the primary unit <strong>of</strong> analysis in most cases.<br />

A notable exception to this generalization is the landscape approach taken on<br />

Garua Island in the Bismarck Archipelago (Torrence & Stevenson 2000). Here the<br />

approach to the landscape sought explicitly to abandon the coastal bias and occupation-<br />

137


site-prospecting objectives <strong>of</strong> previous work in the Bismarcks. In spite <strong>of</strong> a comprehensive<br />

series <strong>of</strong> C14 and obsidian hydration dates on Garua materials <strong>of</strong> roughly Lapita age, the<br />

study faces difficulties associating ages <strong>of</strong> dated materials with the pottery styles found<br />

with them, and the resulting chronological inferences are potentially spreading the span <strong>of</strong><br />

Lapita production/discard rather than refining a ceramic chronology. <strong>The</strong> Garua ceramic<br />

sample would seem to hold considerable potential for a program <strong>of</strong> direct dating <strong>of</strong> local<br />

pottery production styles. <strong>The</strong> focus on inland locations also initially missed significant<br />

Lapita pottery deposits in the sea nearby, but these project-specific problems do not<br />

detract from the value <strong>of</strong> such landscape-based approaches. In Remote Oceania there has<br />

been some effort put into intensive systematic survey <strong>of</strong> landscapes (e.g. Best 1984, Kirch<br />

1988a, Rogers 1973). In the case <strong>of</strong> Niuatoputapu this approach was favoured by<br />

concentrated effort over a number <strong>of</strong> years by two research teams on an island <strong>of</strong> only<br />

15km 2 .<br />

<strong>The</strong> current study seeks method by which the regional distribution <strong>of</strong> Lapita-age<br />

archaeological materials can be established and interpreted in the Roviana<br />

geomorphological setting. While explicit research design in terms <strong>of</strong> sampling theory was<br />

not a feature <strong>of</strong> the 1996 Roviana survey (Sheppard 1996, Sheppard et al. 1999), and data<br />

collected reflect this, both the Roviana and Kaliquongu surveys can pr<strong>of</strong>itably be phrased<br />

in terms <strong>of</strong> sampling theory when assessing the significance <strong>of</strong> results, since they clearly<br />

are samples <strong>of</strong> some sort.<br />

Intensity <strong>of</strong> Survey:<br />

Wobst found in a study <strong>of</strong> archaeologists that they tended to sample a larger landscape less<br />

intensively regardless <strong>of</strong> available resources when asked to design a survey (Wobst<br />

1983:61). A similar tendency is proposed for Near-Oceanian archaeology, although in this<br />

case resources have clearly been limited. Faced with large terrestrial regions, surveys have<br />

tended to try to cover wide regions (see for example Allen et al. 1984) in an opportunistic<br />

manner, relying largely on informant prospection as a site-discovery technique, although<br />

138


this is not universal (as the example <strong>of</strong> Torrence & Stevenson 2000 shows). <strong>The</strong> following<br />

section reviews the regional survey literature in the light <strong>of</strong> Wobst’s observation.<br />

Review <strong>of</strong> Near-oceanic Survey Methods and Results:<br />

Objectives <strong>of</strong> the Review:<br />

One approach to archaeological sampling is to ask how much work needs to be done to<br />

find out what we want to know (Orton 2000:5). By taking a sampling approach to the<br />

Roviana Lagoon early-ceramic survey, the interest is not so much how much survey is<br />

needed, but, given the amount <strong>of</strong> survey that was done, what is the significance <strong>of</strong> the<br />

results towards answering the major research questions? <strong>The</strong> following review <strong>of</strong> Near-<br />

Oceanic survey objectives, methods and results had these aims:<br />

• to critically assess existing knowledge (pertaining to Lapita site survey method,<br />

recorded site density and site preservation/detectability)<br />

• to arrive at a characterization <strong>of</strong> existing knowledge in terms <strong>of</strong> values<br />

(observations) <strong>of</strong> a variable (site density) measured on objects (survey regions)<br />

To do this requires the reworking <strong>of</strong> existing information to arrive at the values <strong>of</strong> this<br />

variable (site density). This is necessarily imprecise in some instances due to the lack <strong>of</strong><br />

specific detail in reports, the need for a definition <strong>of</strong> what constitutes a site, and the unit<br />

<strong>of</strong> measurement (sites per unit area or per unit coastline length).<br />

Method <strong>of</strong> this Review:<br />

<strong>The</strong> literature on archaeological surveys in near-Oceania was examined for information on<br />

survey objectives, methods and results. <strong>The</strong> coastal extent <strong>of</strong> surveys was reconstructed<br />

as far as possible from the literature available. <strong>The</strong> unit <strong>of</strong> measurement used was site<br />

density per km <strong>of</strong> coastline, which poses a problem <strong>of</strong> scale, since coastlines become more<br />

indented at larger map scales. Distances were calculated using published maps <strong>of</strong> survey<br />

139


egions in conjunction with descriptions <strong>of</strong> surveys, using dividers at a scale <strong>of</strong> either 1cm<br />

=1km or 1cm=5km to “walk” the coastline. This method at the latter scale can be expected<br />

to under represent the length <strong>of</strong> indented coastlines relative to the former. <strong>The</strong> effect <strong>of</strong> this<br />

on the conclusions is discussed. Since almost none <strong>of</strong> the studies supply a definition <strong>of</strong><br />

what constitutes a site, structure in the resulting data needs to be discussed in these terms<br />

too. Ideally, we should seek measures <strong>of</strong> intensity <strong>of</strong> use in behavioural measurement<br />

devices such as estimates <strong>of</strong> discarded ceramic vessel quantities rather than site counts, but<br />

there is no data <strong>of</strong> this sort currently available.<br />

At the broader regional scale <strong>of</strong> the Lapita era in near-Oceania, existing knowledge<br />

includes the archaeological record from early colonial missionaries and <strong>of</strong>ficials (e.g. Father<br />

Otto Meyer), Specht’s work at Watom and Buka in the late 60s, the Southeast Solomons<br />

Culture History project and <strong>of</strong>fshoots, Chikamori’s tours, the Lapita Homeland Project and<br />

descendants (e.g. Kirch on Mussau and Summerhayes on Anir), Terrell’s Bougainville<br />

research (Terrell 1976), Irwin’s Shortlands survey, David Roe’s thesis research on<br />

Guadalcanal (Roe 1993), the Solomon Islands National Museum Site recording scheme<br />

(e.g. Miller & Roe 1982, Reeve 1989) (and equivalents in PNG, which have not been<br />

accessed). Of these, some surveys provide little information <strong>of</strong> relevance, and have been<br />

omitted or only briefly discussed. Others are unpublished or difficult to access, so the data<br />

presented below are incomplete, but are data nonetheless. Recent work on Garua is<br />

omitted from this quantitative review, as “site” densities in that case can be expected to far<br />

exceed those recorded elsewhere due to a combination <strong>of</strong> a well preserved buried<br />

landscape <strong>of</strong> Lapita age and a more intensive programme <strong>of</strong> “<strong>of</strong>f-site” testpitting than has<br />

been conducted elsewhere.<br />

What distributional information regarding site density can be extracted from these<br />

field surveys? While none are prepared to consider the results <strong>of</strong> their surveys as unaltered<br />

and contemporaneous settlement patterns, an approach to this problem adopted here is to<br />

look at the range <strong>of</strong> density values for the Lapita period, assuming that many <strong>of</strong> the lower<br />

site densities will reflect low rates <strong>of</strong> site preservation or visibility (together, these create<br />

140


low probability <strong>of</strong> detection), or low-intensity survey coverage/bad luck, and thus low<br />

detection rate rather than absence <strong>of</strong> sites in the past. This leaves a problem in that some<br />

<strong>of</strong> the lower site densities may be an accurate reflection <strong>of</strong> the intensity <strong>of</strong> Lapita<br />

settlement in the past, but discriminating between absent, never present, and absent, not<br />

found is a problem that has to be temporarily sidestepped given the embryonic state <strong>of</strong><br />

existing knowledge. What the data can tell us is roughly what the maximum recorded site<br />

densities are for coarse periods.<br />

In addition to coastline length, a secondary measure <strong>of</strong> survey region area<br />

employed is simply to calculate the area <strong>of</strong> a rectangle on a map roughly encompassing the<br />

area surveyed, subsuming land, sea, coastline length, arable land, reef resources, and any<br />

other factors likely to be significant for site location frequency within this gross<br />

measurement. Density was calculated by dividing the number <strong>of</strong> Lapita sites found by the<br />

number <strong>of</strong> kilometres <strong>of</strong> coastline surveyed, as far as could be determined from the reports.<br />

<strong>The</strong> definition <strong>of</strong> Lapita sites was largely left as per the report, although in the case <strong>of</strong><br />

Nissan (Spriggs 1991) the Halika phase site (Lapita without pots) was not counted as the<br />

review was limited to early ceramic sites where the characteristics <strong>of</strong> the ceramics was the<br />

primary means <strong>of</strong> assignment to the Lapita period. Survey methods were characterized in<br />

terms <strong>of</strong> the following criteria:<br />

• whether survey was predominantly through the informant-prospection method,<br />

expected to create a late-site bias,<br />

• or whether more systematic searching was conducted, expected to yield results<br />

with less <strong>of</strong> a time-bias;<br />

• whether systematic terrestrial pedestrian coastal searches were made, and/or<br />

whether intertidal survey was conducted.<br />

• A further descriptor was added regarding the intensity <strong>of</strong> survey, high, moderate<br />

or low, which was arrived at largely through a subjective assessment <strong>of</strong> details<br />

supplied in the reports on the amount <strong>of</strong> time spent in the survey region.<br />

141


1984 Lapita Homelands Project Survey:<br />

<strong>The</strong> primary aim <strong>of</strong> the 1984 reconnaissance was to explain the aims and nature <strong>of</strong> the<br />

proposed research and to seek reaction to it. A secondary aim was to locate regions to<br />

form base locations for individual projects in the future: there was no aim to carry out full<br />

and extensive surveys <strong>of</strong> such localities (Allen et al. 1984:1). Nevertheless, site surveys<br />

were carried out in Manus, the Mussau Islands and along both coasts <strong>of</strong> New Ireland,<br />

centred at Manggai/Lemakot, Lasigi and Hilalon on the east coast, and Limpos on the west<br />

coast. Site surveys were carried out also in the Jaquinot Bay area <strong>of</strong> East New Britain<br />

(Allen et al. 1984:3,6-7).<br />

Specht and Allen spent seven days in the Mussau area, searching principally the<br />

small islands <strong>of</strong> Eloaua and Emananus. Fourteen <strong>of</strong> sixteen new sites were recorded on<br />

these small islands. Of the two sites on the main island, one was reported by the provincial<br />

member, while the other was spotted during a brief visit to Lomakuauru village for<br />

purposes <strong>of</strong> consultation. <strong>The</strong>se fourteen sites on the small islands comprised Lapita<br />

middens, other or undiagnostic pottery-bearing middens, and aceramic middens. This seven<br />

day survey <strong>of</strong> roughly thirty square kilometres (land and sea, my calculation) must rate as<br />

one <strong>of</strong> the more intensive in Near-Oceania. An approximate measurement <strong>of</strong> coastline<br />

length from the published map (Kirch’s section in Allen et al. 1984:146) yielded a<br />

measurement <strong>of</strong> thirty-three kilometres, or a recorded Lapita site density <strong>of</strong> 0.09 sites per<br />

km.<br />

<strong>The</strong> 1984 New Ireland survey was more in the nature <strong>of</strong> a tour, with heavy<br />

emphasis on cave sites. <strong>The</strong> Lemau pottery-bearing mounds are assumed to have been<br />

discovered in passing by road, rather than by pedestrian survey, attention being drawn to<br />

these high-visibility features by the favourable location <strong>of</strong> this village on a deepwater bay<br />

(Allen et al. 1984:17-18).<br />

Specht, Ambrose and Allen spent five days in the Pomio-Palamal region <strong>of</strong><br />

Jaquinot Bay, and found the area poor in archaeological resources despite good shelter in<br />

the bay from monsoon-season seas. No pottery sherds were seen in this time. <strong>The</strong><br />

142


possibility <strong>of</strong> poor preservation as a result <strong>of</strong> high local rainfall was noted (5-6000mm per<br />

annum). <strong>The</strong>ir assessment <strong>of</strong> the area as potentially having a low level <strong>of</strong> detectability<br />

means this negative result should be disregarded for the present purposes.<br />

<strong>The</strong> Talasea area was not re-surveyed in 1984, as Specht had surveyed previously<br />

(Specht 1974). Specht and Kamminga had surveyed a 60km stretch <strong>of</strong> coast on the north<br />

coast <strong>of</strong> the Huon peninsula in 1972, without finding Lapita pottery or obsidian. <strong>The</strong><br />

discovery in 1973 at Talasea <strong>of</strong> nine coastal ceramic sites, including three identifiable<br />

Lapita sites was therefore notable at the time, as this seemed to Specht to add weight to<br />

an interpretation <strong>of</strong> an Oceanic distribution <strong>of</strong> Lapita, extending close to, but not to, the<br />

north New Guinea coast.<br />

Specht notes a total <strong>of</strong> 15-20 sites with some dentate sherds, but only three or four<br />

sites with large quantities <strong>of</strong> dentate sherds, in a total coastal survey region <strong>of</strong> 25km,<br />

surveyed by a mix <strong>of</strong> intensive pedestrian cover and informant prospection (the site at<br />

Pasiloke was found by the latter method) (Specht, pers. comm. 2001). Coastal Talasea<br />

Lapita sites were located in some cases in the intertidal (for example the FDK obsidian<br />

source and Lapita findspot (Specht 1981), and in others on the elevated reef shore platform<br />

slightly above sea level, and eroding into the sea. Major disturbance by bulldozers for<br />

roading gravel seems to have been a factor in the discovery <strong>of</strong> at least site FCS (Specht<br />

1974). Four sites in 25km yields a Lapita distribution <strong>of</strong> 0.16 sites per km, or if the larger<br />

total is used, the site density is 20 ÷ 25 = 0.8 sites per km.<br />

<strong>The</strong> 1984 party revisited previously reported <strong>of</strong>fshore sites at Boduna and Garala<br />

islands, and mention water-rolled pottery above the high water mark on islands then<br />

thought to be sinking, suggested either a complex recent sea-level history or storm<br />

deposits relocating materials from now-submerged sites. Another explanation not<br />

considered in the report might be an intertidal distribution <strong>of</strong> settlement in the past. Large,<br />

volcanic-tempered sherds were present in the sea at Boduna and Garala islands, and were<br />

notable for their strength compared to the s<strong>of</strong>t Mussau sherds which in contrast were<br />

exposed on land and calcareous tempered (Allen 1984:19-20, Ambrose & Gosden 1991).<br />

143


Recent collection <strong>of</strong> ceramics from the reef flat on Boduna (Torrence & White 2002,<br />

White et al. 2002), separated from Talasea by a 1km channel, heightens the likelihood <strong>of</strong><br />

an intertidal component <strong>of</strong> the archaeological distribution in the Talasea area, rather like<br />

the situation at Buka, where both terrestrial and intertidal ceramic sites are found.<br />

It has been pointed out that there seems to be ample evidence that Lapita<br />

settlement pattern at Talasea differs from the stilt-house patterning noted by Gosden for<br />

the Arawe Islands (Specht et al. 1991). Specht notes evidence for massive landscape<br />

changes in the Willaumez region, and also suggests that high site density is related to<br />

intensity <strong>of</strong> survey and monitoring over an extended period (Specht, pers. comm. 2001).<br />

Lapita Homelands Project, 1985 Season:<br />

While the 1985 season was reported in more final form in 1991 (Allen & Gosden 1991),<br />

some significant additional information regarding survey is contained in the 1985 field<br />

season report. In relation to work at Bal<strong>of</strong>,<br />

“Explorations for coastal sites were also carried out. Three were found<br />

(ELA, ELB, ELD), each with small obsidian flakes and red, shell tempered<br />

pottery. At ELD two crenulate rims, three sherds with applied nubbins,<br />

some incised decoration and applied strips all link this pottery with that<br />

previously found at Lesu. However, all sites had been very disturbed, and<br />

no undisturbed deposits were found (White et al. 1985)”<br />

No details were given <strong>of</strong> the extent or type <strong>of</strong> survey which located these coastal sites, and<br />

they make no appearance in the 1991 volume, although presumably do exist in the PNG<br />

museum records.<br />

Kirch’s group conducted further survey on the smaller islets near Eloaua, locating<br />

seven new sites, principally aceramic middens (Kirch 1985). Kirch and Gorecki visited<br />

Lavongai, New Hanover, for two days and surveyed beaches, river banks and valleys in<br />

the area (Gorecki 1985). Middens and low mounds were noted along virtually the entire<br />

Lavongai bay, with no pottery found, although the National Museum had previously<br />

received some sherds from there. One <strong>of</strong> the research questions for the 1988 Mussau field<br />

season was to define the chronology and nature <strong>of</strong> the transition from Lapita to post-<br />

144


Lapita, involving excavation <strong>of</strong> an additional aceramic midden site, but no further survey<br />

seems to have occurred by the termination <strong>of</strong> the field phase <strong>of</strong> the project (Kirch 2001:10-<br />

24)<br />

<strong>The</strong>re is considerably more detail on survey method in the Duke <strong>of</strong> Yorks Islands<br />

in Lilley’s contribution to the 1985 field season report (Lilley 1985) than in the subsequent<br />

1991 volume. Intensive investigation <strong>of</strong> the western coasts <strong>of</strong> Duke <strong>of</strong> York, Makada and<br />

Mioko Islands located six Lapita sites. Inland survey transects fostered the opinion that this<br />

was a zone <strong>of</strong> low archaeological potential, and the eastern coastlines were cliffed or<br />

swampy and considered unsuitable for settlement, or possibly having poor archaeological<br />

visibility: this point is not clear. Makada, Duke <strong>of</strong> York and Mioko Islands were surveyed,<br />

predominantly along favourable sections <strong>of</strong> the coastlines for settlement (Lilley 1991). <strong>The</strong><br />

survey comprised a total land and sea area <strong>of</strong> about 75km 2 , yielding a Lapita site density<br />

<strong>of</strong> 0.08 sites/km 2 . This included about 28km <strong>of</strong> coastline, yielding a site density <strong>of</strong> 0.214<br />

sites per km. A subsequent 5-week intensive pedestrian survey <strong>of</strong> 60km <strong>of</strong> coastline (my<br />

measurement) located well-preserved Lapita potttery below the water table at sites SDQ<br />

and SDP, site SEE was found to have a high density <strong>of</strong> Lapita sherds, and SEP had some<br />

Lapita sherds in the lowest Layers (White & Harris 1997). <strong>The</strong> more exclusive definition<br />

<strong>of</strong> what constitutes a site, compared to the earlier survey, means that site dentity decreases<br />

compared to Lilley’s survey, to 67 sites per 1000km, or 0.025 sites per km 2 (I measured<br />

total area, land and sea at 160km 2 ).<br />

In the Arawe Islands, Specht spent five nights on Pililo, and three surface pottery<br />

sites/findspots were recorded, two with Lapita sherds (Specht 1985). On Kumbun, local<br />

residents identified three ceramic sites, Pukwo site with Madang/Sio-Gitua pottery, Iolmo<br />

cave with similar pottery, and Emol beach flat, with similar pottery again. Arawe Island<br />

was visited briefly, and Lapita was immediately discovered together with Madang/Sio<br />

Gitua, and other styles, on the Makegur spit. Specht left at this point, leaving Gosden to<br />

continue the survey (Gosden 1985).<br />

After an unrewarding search for in situ stratified deposits in the Arawe island site,<br />

145


Gosden spent some days excavating in Paligmete village on Pililo island, then “surveyed<br />

the inhabited islands”, revisiting Arawe and Kumbun, and briefly visiting Maklo and<br />

Kauptimete, where an extensive midden with obsidian and a single Lapita sherd was<br />

discovered. A similar site with unspecified pottery was discovered on Maklo. Gosden<br />

describes these surveys as “by no means comprehensive”. Agussak island was visited, and<br />

a pottery site with one possible Lapita sherd was recorded. Seventeen sites (six Lapita)<br />

were recorded on six islands totaling 110 km 2 approximately (land and sea, my calculation<br />

from the details given in the 1985 and 1991 reports).<br />

Gosden finally reported a cluster <strong>of</strong> seven Lapita sites within an area <strong>of</strong> around<br />

100km 2 (land and sea) in the Kauptimete/MakIo/Kumbun/Adwe/Pililo region <strong>of</strong> the Arawe<br />

area, although he was not specific about the extent and method <strong>of</strong> survey. This yielded a<br />

site density <strong>of</strong> roughly 0.07 sites per km 2 (Gosden 1991b). Gosden questioned whether<br />

these sites were contemporaneous or not, and expected other sites would be found in the<br />

future. In spite <strong>of</strong> this caveat, he subsequently suggests the clustered distribution<br />

comprised a social landscape (Gosden & Webb 1994). My coastline measurement yielded<br />

a total <strong>of</strong> 42km for the islands surveyed, or 0.17 sites per square kilometre. It seems sites<br />

with small amounts <strong>of</strong> Lapita pottery are included in these totals.<br />

In the Kandrian district, Specht revisited terrestrial Lapita find spots on Apugi<br />

island and conducted test excavations, without any survey coverage details beyond the<br />

level <strong>of</strong> village names (Specht 1985). During this fieldwork he received insistent reports<br />

<strong>of</strong> pottery on the reef at Kreslo (Specht 1985, 1991) and spent a day visiting the site. His<br />

initial conclusion was that this was a defluved beach deposit but on reflection and in view<br />

<strong>of</strong> results since obtained on Mussau and in the Arawes, Specht noted the difficulty posed<br />

to this explanation by the absence <strong>of</strong> sherds on the adjacent land, and suggested stilt houses<br />

as a possible site-formation process (Specht 1991).<br />

For the seven hundred square kilometres or so <strong>of</strong> the Siassi Islands, a single Lapita<br />

site on Tuam (Lilley 1986: 109&166) gave a rough site density <strong>of</strong> 0.0014 sites/km 2 , or<br />

using my coastline measurement <strong>of</strong> 130km (probably an underestimate due to a crenellate<br />

146


coastline drawn at a smaller scale than the others) 0.008 sites per km. Lilley noted coastline<br />

subsidence and emergence (Lilley 1986:105) and the effects <strong>of</strong> the 1888 Ritter tsunami and<br />

storm action on this coastline. More intensive coastal survey was conducted on the small<br />

islands <strong>of</strong> Tuam and Malai, with a combined coastline <strong>of</strong> around 10km, and along a 10km<br />

stretch <strong>of</strong> the northeast coast. This yielded a much higher return <strong>of</strong> one Lapita site for 10-<br />

20 km <strong>of</strong> coast, depending on how one views the preservation potential <strong>of</strong> the northwest<br />

coast, giving a figure <strong>of</strong> 0.1 to 0.05 Lapita sites per km.<br />

On Nissan, three or four Lapita sites were discovered (Spriggs 1991), depending<br />

on whether Halika-phase sites can be regarded as “Lapita without pots”. Spriggs notes that<br />

“Systematic surveys <strong>of</strong> former village sites were not undertaken, although<br />

coverage was good for the south part <strong>of</strong> Nissan. Most surface sites from<br />

Tanamalit to Nahoi were destroyed by construction <strong>of</strong> a large American<br />

base in WWII, as were some areas <strong>of</strong> Barahun island and Iachtibol.<br />

Nachman had collected pottery from 105 named localities in the vicinity <strong>of</strong><br />

Balil, Siar, Salepen and Poriwan villages and on Sirot island, and so no<br />

resurvey <strong>of</strong> these areas for village sites was undertaken, although many<br />

such sites were noted in passing on Sirot and Northwest Nissan.” (Spriggs<br />

1991:226)<br />

Spriggs’ survey had a strong emphasis on rockshelters, with 45 rockshelters and 21 open<br />

sites recorded (Spriggs 1985). Of these, the reef site at Tarmon village was the only open<br />

Lapita site, the other two being rockshelter deposits with small samples <strong>of</strong> sherds<br />

recovered. <strong>The</strong> atoll covers an area <strong>of</strong> about 100 square kilometres, yielding an open-site<br />

density <strong>of</strong> 0.01 sites per square kilometre, or for an estimated 34 km <strong>of</strong> lagoon-side<br />

shoreline (70km total including ocean-side shoreline) 0.029 sites per km.<br />

Southeast Solomon Islands:<br />

Survey and excavation on Santa Ana (Davenport 1972) located pottery at Feru, leading<br />

Davenport to hypothesize an early, Lapita-related ceramic settlement, but no information<br />

on site distribution can be derived from the report. Survey method is not reported in any<br />

detail, and is assumed by default to comprise a tour <strong>of</strong> villages seeking local information,<br />

with opportunistic examination <strong>of</strong> areas en route, with prospection rather than<br />

147


distributional information as the primary objective.<br />

On Bellona, Poulsen reported a number <strong>of</strong> terrestrial sites sealed by earth mounds,<br />

and demonstrated that the Sikumango mound sealed deposits that contained a sherd that<br />

“looked like Lapita” (Poulsen & Polach 1972). It is clear from Poulsen’s account and<br />

radiocarbon dates that site visibility for this type <strong>of</strong> deposit may bear little relation to the<br />

distribution <strong>of</strong> earth mounds, and therefore the 33km 2 Bellona survey (land and sea) can<br />

only tell us that there was at least one (probable) Lapita site there. Survey methods seem<br />

to have been more intensive than the subsequent reconnaissance efforts <strong>of</strong> others elsewhere<br />

in the Southeast Solomons, with a month spent covering the 9km 2 or so <strong>of</strong> the inland<br />

garden zone and limited coastal area suitable for settlement. Survey seemed principally<br />

aimed at mapping the conspicuous mound and pit features, abandoned pre-WWII<br />

habitation sites, caves in the bush, and freshwater sources on the coast; the distribution <strong>of</strong><br />

which were recorded in detail. <strong>The</strong> results <strong>of</strong> Poulsen’s excavations during the following<br />

month raise the possibility <strong>of</strong> an early Lapita phase <strong>of</strong> low archaeological visibility, that<br />

may bear little distributional relationship beyond general environmental constraints to the<br />

mapped distribution <strong>of</strong> mound sites, which seem to have formed partly through occupation<br />

during the period 480-1290 AD, approximately.<br />

Southeast Solomon Islands Culture History Project:<br />

Green conducted a week-long terrestrial survey <strong>of</strong> the Surville peninsula on Makira (San<br />

Cristobal) in 1970 (Green 1976c), with the aim, “...to determine the types and range in<br />

sites in this area.” <strong>The</strong> survey encompassed an area <strong>of</strong> 250km 2 (my calculation, land and<br />

sea) or about 60km <strong>of</strong> coastline (measured from map presented in Green 1976c), and in<br />

spite <strong>of</strong> good surface visibility <strong>of</strong> artefacts as a result <strong>of</strong> crab burrowing, and a relatively<br />

rich archaeological landscape <strong>of</strong> late-prehistoric coastal terrestrial sites, no early ceramics<br />

were found. Methods are assumed to have been predominantly informant-prospection.<br />

Green regarded the area as favourable for settlement, with good canoe access to a<br />

relatively sheltered harbour, and the complete lack <strong>of</strong> any pottery must have been food for<br />

148


thought, particularly in comparison to the record in the Reef/Santa-Cruz group. Green<br />

noted the dramatic effects <strong>of</strong> 1971 cyclone waves on coastal rockshelters, a comment<br />

which now seems relevant to the question <strong>of</strong> intertidal site preservation on Makira. While<br />

this survey is robust evidence for the absence or rarity <strong>of</strong> terrestrial Lapita pottery sites,<br />

it cannot tell us whether Lapita was deposited in the sea in the past from stilt villages.<br />

A series <strong>of</strong> surveys <strong>of</strong> Ulawa (Hendren 1976, Ward 1976) covered an approximate<br />

area (land and sea) <strong>of</strong> 209km 2 (my calculation) and revealed a coastal terrestrial focus <strong>of</strong><br />

late prehistoric settlement. No details <strong>of</strong> prospection/survey method are reported. Hendren<br />

noted an almost continuous distribution <strong>of</strong> tool and midden scatter in coastal areas. Ward<br />

notes evidence for exposure <strong>of</strong> this coast to cyclone waves, and effects on terrestrial site<br />

preservation. <strong>The</strong> occupation sites excavated by Ward postdated about 700AD. Results<br />

suggest terrestrial pottery sites <strong>of</strong> any period are rare or absent from the extant<br />

archaeological record.<br />

Green’s 1970 reconnaissance survey <strong>of</strong> Uki located sites from the last 500 years<br />

on a series <strong>of</strong> uplifted shorelines (Green 1976b). No pottery was found. In the present<br />

context these results raise the question whether survey methods were sufficiently intense<br />

to locate low-visibility ceramic deposits had these been formerly located in the intertidal<br />

areas presently forming a back-swamp adjacent to the paleoshorelines.<br />

Inland survey <strong>of</strong> Santa Cruz (Nendo) (Yen 1976) located late prehistoric or historic<br />

occupation sites and no pottery.<br />

Green found six Lapita sites over the 78 km 2 (land and sea?) <strong>of</strong> the main Reef<br />

Islands (although in remote-Oceania, included for comparative purposes), and two sites<br />

on the roughly 660 km 2 <strong>of</strong> Santa Cruz. Green reported extensive searching and questioning<br />

<strong>of</strong> local people, without further result (Green 1976a), from which I calculated a range <strong>of</strong><br />

site densities from 0.077 to 0.00303 sites/km 2 . By my measurements, the main Reef<br />

Islands, and Santa Cruz have coastlines <strong>of</strong> 33 and120km (eastern portion as per Green<br />

p249) respectively, giving site densities <strong>of</strong> 0.182 sites per km <strong>of</strong> coast for the Reefs, and<br />

0.012 sites per km for Santa Cruz.<br />

149


A subsequent brief resurvey <strong>of</strong> the western end <strong>of</strong> Santa Cruz, the Graciosa Bay<br />

area (Mccoy & Cleghorn 1988:106) made heavy use <strong>of</strong> informant-prospection. <strong>The</strong> main<br />

emphasis <strong>of</strong> the fieldwork was on area excavations. A two-day survey <strong>of</strong> the coastline <strong>of</strong><br />

To Motu Neo (about 20km from the map published by McCoy and Cleghorn) located an<br />

additional three sites, bringing the To Motu Neo total to five. <strong>The</strong>re is no mention <strong>of</strong><br />

Lapita at the two unexcavated rockshelters, but the Novlao Rockshelter site yielded plain<br />

ceramics and Lapita diagnostic decoration, and C14 dates spanning the Lapita period. <strong>The</strong><br />

intensive 20km coastal survey thus located two Lapita open sites without structural<br />

features, and one Lapita rockshelter/open site with structural features. Site density works<br />

out at 0.15 sites per km <strong>of</strong> coastline for an intensive survey method.<br />

Central Solomon Islands:<br />

Roe’s surveys on Guadalcanal had the following aims:<br />

• to identify the range <strong>of</strong> site types known to local informants as a key to the late<br />

prehistoric/historic settlement patterns<br />

• prospection, to locate archaeological sites with excavation potential<br />

• to search specifically for rock art sites and inland agricultural sites<br />

(Roe 1993: 31)<br />

In forested areas with dissected limestone geology, local informants were the<br />

dominant means <strong>of</strong> prospection for rockshelter sites. In the open grassland <strong>of</strong> the<br />

Northwest coast regional survey, site survey was conducted without prior local<br />

information, but no further information on method was given (Roe 1993:31-32). No<br />

pottery was seen during these surveys, and no sites <strong>of</strong> any kind were found during the<br />

north coast survey. Roe suggested that construction <strong>of</strong> the North Guadalcanal coastal<br />

road, clearance for plantations, and large-scale WWII bulldozing may have been major<br />

factors in creating this archaeological blank but didn’t consider the effects <strong>of</strong> cyclone storm<br />

surge or tsunami on site preservation along this coast.<br />

In a review <strong>of</strong> Solomon Islands prehistory prior to his research, which was<br />

150


considered suggestive <strong>of</strong> a different cultural history than that recorded elsewhere, due to<br />

the absence <strong>of</strong> pottery in surface collections, Roe included a footnote to the effect that<br />

Spriggs’ work on Nissan revealing Lapita occupation, Wickler’s Buka work that found<br />

Lapita on the reef flats, and Reeve’s data on the intertidal ceramics in the Western<br />

Province were all unknown to him at the time <strong>of</strong> fieldwork (Roe 1993:6, 10). What Roe’s<br />

work does tell us is that there is a contrast in the terrestrial record between Guadalcanal<br />

and the north Solomon Islands in the absence <strong>of</strong> late-prehistoric pottery, but absence <strong>of</strong><br />

evidence regarding the intertidal/reef flat record means that this contrast cannot be safely<br />

extended back to the Lapita period given current knowledge. It seems clear enough that<br />

Roe did not set out to find the type <strong>of</strong> reef-flat evidence found on Buka and Nissan, and<br />

that the absence <strong>of</strong> such evidence in his results cannot be taken as evidence for absence.<br />

Roe’s results therefore are limited to negative locational evidence, which in itself is weak<br />

due to recent coastal landscape alterations in the area surveyed.<br />

Cultural Resource Management (CRM) surveys by the National Museum <strong>of</strong> Solomon<br />

Islands:<br />

<strong>The</strong> Solomon Islands National Sites Survey listed among its objectives the assessment <strong>of</strong><br />

archaeological potential <strong>of</strong> each island (Miller & Roe 1982). This is a similar objective to<br />

many <strong>of</strong> the surveys conducted within the Lapita Homeland Project. It appears that the<br />

larger regional surveys were all predominantly reliant on the informant method <strong>of</strong><br />

prospection. Of these only a survey <strong>of</strong> Simbo located pottery deposits. <strong>The</strong>se surveys<br />

included the southeast Choiseul survey, the survey <strong>of</strong> the Bughotu area on Isabel, the<br />

Kwaio area survey on Malaita, a survey <strong>of</strong> Simbo (all prior to 1978); Paripao district,<br />

inland Guadalcanal, Santa Catalina survey, and survey <strong>of</strong> Dorio coastal district on Malaita<br />

(between 1978 and 1980).<br />

Within the National Sites Survey, survey in advance <strong>of</strong> construction or logging has been<br />

more likely to turn up pottery deposits, possibly due partly to more intense and unbiased<br />

151


survey methods. Such surveys include four surveys on Kolombangara, all <strong>of</strong> which located<br />

pottery, coastal plain survey in the Arosi district on Makira, which located no pottery,<br />

survey in the Viru harbour area <strong>of</strong> New Georgia which located a range <strong>of</strong> inland sites but<br />

no pottery (given the results <strong>of</strong> the NGAS project, its absence is anomalous). Survey in<br />

advance <strong>of</strong> mining on Vaghena resulted in test-excavations on pottery-bearing sites. Survey<br />

in advance <strong>of</strong> logging on north New-Georgia recorded significant pottery-bearing sites.<br />

Survey on the north coast <strong>of</strong> Nggela in advance <strong>of</strong> road construction located no pottery<br />

sites, but a range <strong>of</strong> inland sites. Survey on the Shortland islands in advance <strong>of</strong> logging<br />

recorded pottery sites. Miller and Roe concluded that:<br />

“Pottery, previously regarded as being restricted in distribution to the<br />

south-east Solomons, northern Choiseul and the Shortland Islands, appears<br />

to have been in use at some time throughout the greater part <strong>of</strong> the western<br />

Solomons (Miller & Roe 1982).”<br />

None <strong>of</strong> the sites located in these surveys can be described as Lapita, although Reeve<br />

characterized the intertidal Roviana pottery seen later at Paniavile and at other unspecified<br />

locations as possibly Lapita-derived.<br />

<strong>The</strong>se surveys yield no information on the distribution <strong>of</strong> Lapita sites in near-<br />

Oceania, other than some negative evidence against a terrestrial distribution. As for Roe’s<br />

work on Guadalcanal, in view <strong>of</strong> what seem to be many examples <strong>of</strong> intertidal Lapita<br />

settlements across the northern part <strong>of</strong> the region at present, as for many <strong>of</strong> the researchoriented<br />

surveys, these CRM surveys <strong>of</strong>fer no strong evidence for absence <strong>of</strong> Lapita in<br />

southern near-Oceania, but may be evidence for a relative rarity <strong>of</strong> Lapita sites, or less<br />

visible locational patterning than in the Northern region.<br />

A number <strong>of</strong> more recent CRM survey reports are held at the National Museum<br />

and in provincial <strong>of</strong>fices, but were not accessible at the time <strong>of</strong> writing.<br />

Bougainville and the Shortland Islands:<br />

A 1966 survey report (Lampert 1966) was not accessed. In the Buka area, Specht<br />

conducted an extensive coastal survey (Specht 1969), while Wickler restricted his survey<br />

to coastal areas suitable for settlement (Wickler 1995:62-68). Wickler’s survey yielded<br />

three Lapita sites on the reef flats, in some cases extending (as separate sites) on to shore,<br />

152


and at least one other non-cave site (DAI) (Wickler 1995: 714-16). <strong>The</strong> Lapita site density<br />

for the approximately 600km 2 region works out to 0.007 sites/km 2 .<br />

Wickler’s survey aims were to locate sites suitable for excavation rather than to<br />

achieve extensive areal coverage (Wickler 2001:63). Survey <strong>of</strong> the east coast <strong>of</strong> Buka was<br />

by vehicle, with likely locations on raised coral flats, beaches and reef flats examined. In<br />

common with Specht and Terrell (for the Silao Peninsula), Wickler noted the virtually<br />

continuous distribution <strong>of</strong> archaeological materials along the raised coral portions <strong>of</strong> this<br />

shoreline, and an arbitrary classification into sites. It would appear from the detail given<br />

(Wickler 2001:63-68, 215-216) that around 90km was covered in this manner. For the<br />

Lapita period, three early-phase sites were found on reef flats, and two open terrestrial<br />

sites. For Wickler’s more intensive survey <strong>of</strong> favourable locations these results yield a site<br />

density <strong>of</strong> 0.056 Lapita sites per km <strong>of</strong> coastline surveyed (total km rather than coastline<br />

length <strong>of</strong> targeted locations).<br />

A surface collection <strong>of</strong> sherds from Teop was reported in 1964 (Shutler & Shutler<br />

1964). No details <strong>of</strong> this were available to the author. Terrell surveyed four regions on<br />

Bougainville, Silao Area, Teop area and Numa Numa area on the east coast in<br />

north/central Bougainville, and Paubake area <strong>of</strong> south Bougainville around Buin. Terrell<br />

noted the pioneering exploratory nature <strong>of</strong> his surveys, and the extensive rather than<br />

intensive approach (Terrell 1976:218). He wanted to sample extensively to find generalities<br />

rather than the specific sequence <strong>of</strong> a local area, and wanted to include both Austronesian<br />

and Papuan language areas in his sample (Terrell 1976:224-227). Terrell conducted a<br />

second survey at Buin following the initial extensive reconnaissance, undertaken in view<br />

<strong>of</strong> the divergent results obtained for this locality on the initial visit.<br />

Methods included informant-prospection, and can be expected to have yielded<br />

results strongly biased toward the late-prehistoric and historic settlement pattern. Site<br />

survey on the Silao peninsula revealed the northern coast <strong>of</strong> the peninsula to be “one long<br />

archaeological site” for the approximately 14 km <strong>of</strong> coastline surveyed, echoing Specht’s<br />

description. Fifty-two “sites” were abstracted from this near continuous distribution, from<br />

which almost 100 locality collections were made. No sherds in the Buka (Lapita) style<br />

were identified, but the low number <strong>of</strong> these in Specht’s samples was considered by<br />

153


Terrell to make sampling error a sufficient explanation for their absence. <strong>The</strong> pottery all<br />

fell within Specht’s Sohano and later sequence.<br />

In the Teop area, approximately 19km <strong>of</strong> coastline was surveyed, as well as some<br />

inland journeying. Much <strong>of</strong> the coast was either mangrove flats or swampy, and only four<br />

prehistoric pottery sites were recorded on the upraised coral terrestrial coastline, these<br />

being styles from Specht’s sequence, back to at least the Sohano style. With the exception<br />

<strong>of</strong> the Nanava form, thought to have been manufactured to the south near Kieta, these<br />

were locally produced. (Terrell 1976:240). Sherds were scarce in the interior.<br />

In the Numa Numa area, approximately 45km <strong>of</strong> coastline were surveyed, locating<br />

32 sites in spite <strong>of</strong> low visibility caused by dense coastal bush along much <strong>of</strong> this sparsely-<br />

occupied coast. Mararing to recent styles were found (Terrell 1976: 252-257). Terrell<br />

considered these surveys too preliminary and incomplete to assess temporal changes in site<br />

density.<br />

<strong>The</strong> Paubake survey area in the south <strong>of</strong> Bougainville seemed on prior information<br />

to contain a ceramic sequence that differed from Specht’s Buka sequence (Terrell<br />

1976:261-263, Figure 4.7). About 8km <strong>of</strong> coastline was included in the survey. Sixteen<br />

sites were initially recorded, including surface pottery scatters and stone constructions,<br />

with sherds present in smaller numbers than the northern surveys. An additional 44 sites<br />

were visited and mapped during the second survey in this region. Only one <strong>of</strong> these sites<br />

was located on the coast,and it is not reported whether this can be considered to represent<br />

a distributional feature <strong>of</strong> past settlement. No Lapita pottery was seen.<br />

Terrell’s surveys in these four areas are incomplete, in his own estimation, and<br />

cannot be taken as strong evidence for the absence <strong>of</strong> Lapita occupation in the past. All<br />

that can be said with confidence is that nowhere in the areas surveyed is Lapita pottery an<br />

obvious feature <strong>of</strong> the coastal terrestrial archaeological record. We can only speculate as<br />

to whether this is to do with past behaviour patterns that contrast with areas where Lapita<br />

pottery is a more obvious component <strong>of</strong> the record, (this could include more subtle<br />

behavioural variants than avoidance <strong>of</strong> the area by Lapita-using people, e.g. intertidal site<br />

location in the past) or whether site visibility and/or preservation factors can account for<br />

such variability in the Lapita archaeological record. Given Wickler’s survey results, I<br />

would expect that Lapita sites were there, in the sea, 3000 years ago.<br />

154


<strong>The</strong> coastal portions <strong>of</strong> Terrell’s surveys were all terrestrial, and total<br />

approximately 86km <strong>of</strong> coastline on Bougainville, surveyed in a low-intensity manner,<br />

using predominantly informant-prospection and, for much <strong>of</strong> his sample, having low<br />

archaeological visibility due to vegetation cover. All the sites recorded belonged to the last<br />

millennium. As noted above, such survey methods can be expected to create a strong late-<br />

site bias.<br />

Irwin’s survey <strong>of</strong> the Shortland Islands began by surveying<br />

“a large area in order to get some idea <strong>of</strong> the nature and variation <strong>of</strong> the surface evidence<br />

in addition to its distribution.” (Irwin 1972:3)<br />

Table 5: Summary <strong>of</strong> a review <strong>of</strong> survey methods and Lapita results.<br />

SURVEY REGION SURVEY METHOD SURVEY INTENSITY COASTLINE<br />

SURVEYED (KM)<br />

Eloaua/ Emananus<br />

1984<br />

terrestrial<br />

?<br />

Huon Pen. ? high but unfavourable for<br />

settlement (specht pers comm 2001)<br />

155<br />

LAPITA<br />

OPEN SITES<br />

high 33 3 91<br />

60 0 0<br />

DENSITY<br />

(SITES /<br />

1000KM)<br />

Talasea ‘73 terrestrial intertidal high? 25 4-20 160-800<br />

Duke <strong>of</strong> York terrestrial<br />

?<br />

Moderate 28 6 214<br />

Arawe terrestrial moderate 42 7 167<br />

Siassi terrestrial high 10 1 100<br />

Nissan terrestrial<br />

intertidal?<br />

late bias for open sites<br />

moderate 34 1 29<br />

Bellona terrestrial, late bias high - 0 0<br />

Makira terrestrial ? not meas 0 0<br />

Ulawa late bias low 0 0<br />

Uki paleo-shoreline ? 0 0<br />

Santa Cruz terrestrial<br />

inf. prosp.<br />

Reef Islands terrestrial<br />

inf. prosp.<br />

Guadalcanal terrestrial<br />

varies<br />

low to mod. 120 2 12<br />

high 33 6 182<br />

modified coastal landscape ? 0 0<br />

Buka (Wickler) terrestrial intertidal moderate 90 5 56<br />

Bougainville terrestrial<br />

inf. prosp.<br />

Shortland terrestrial<br />

inf. prosp.<br />

Low 86 0 0<br />

targeted 70 0 0<br />

Duke <strong>of</strong> Yorks Terrestrial intensive 60 4 67<br />

To Motu Neo Terrestrial intensive 20 3 150


Survey was concentrated on the east and south coasts <strong>of</strong> Alu Island, and appears to have<br />

comprised a mix <strong>of</strong> informant-prospection and highly targeted terrestrial coverage <strong>of</strong> the<br />

about two-thirds <strong>of</strong> the lengthy Alu coastline, within an overall survey block <strong>of</strong> about<br />

500km 2 . A rough estimate <strong>of</strong> the length <strong>of</strong> coastline covered is 70km, <strong>of</strong> a similar<br />

magnitude to the total coastal portion <strong>of</strong> Terrell’s Bougainville survey. His approach was<br />

highly targeted towards reef-passage locations (Irwin 1972:19), and made use <strong>of</strong><br />

informant-prospection method. Twenty-two sites were recorded, most <strong>of</strong> which yielded<br />

surface pottery collections, which were fitted into a hypothetical post-500AD sequence<br />

comprising two widely-spaced early-period sites, nine closely spaced middle-period sites,<br />

and five closely spaced late-period sites (Irwin 1972:193-196). No Lapita sites were found.<br />

Irwin’s survey seems to provide reasonable evidence that terrestrial Lapita sites were either<br />

rare or absent at the time <strong>of</strong> the survey, and likely to have been so in the past. No strong<br />

evidence for the absence <strong>of</strong> either terrestrial or intertidal Lapita can be claimed though. <strong>The</strong><br />

metric results <strong>of</strong> this analysis <strong>of</strong> survey method and results in Near Oceania and the<br />

Southeast Solomons are summarized in table Table 5.<br />

Discussion <strong>of</strong> Review Results:<br />

Many have commented on the gross patterning evident from these survey results, which<br />

is the absence <strong>of</strong> Lapita pottery in near-Oceania south <strong>of</strong> Buka, and on the Continental<br />

New Guinea coast (Huon Peninsula). <strong>The</strong>se broad-scale discussions (see for example<br />

Anderson 2002) have taken place within a data framework with minimal information on<br />

or discussion <strong>of</strong> site preservation and visibility, despite comments within the individual<br />

156


survey reports on such factors, for example on differential exposure to cyclone and<br />

tsunami waves. With an average return for these 17 surveys <strong>of</strong> 2-3 Lapita sites each,<br />

absence <strong>of</strong> sites through sampling error alone is a distinct possibility, even without<br />

consideration <strong>of</strong> survey coverage, site preservation and site visibility. None <strong>of</strong> the surveys<br />

in this review which returned a negative result for Lapita can be viewed as evidence for<br />

absence, when survey methods and preservation/visibility are taken into account, if one<br />

allows that settlement in the past may have been over water in the Lapita period for the<br />

survey regions which returned negative results. It must also be remembered that there is<br />

very little information about relative sea-level changes in most <strong>of</strong> the surveys.<br />

For Near-Oceania, in regions where Lapita occupation has been demonstrated,<br />

Lapita site density ranged from 29 to 800 sites per 1000km <strong>of</strong> shoreline, with an average<br />

density <strong>of</strong> 112-183 sites per 1000km, depending on the definition <strong>of</strong> a Lapita site. <strong>The</strong><br />

sample <strong>of</strong> surveys includes two remote-Oceanic surveys, and seven near-Oceanic surveys.<br />

Much <strong>of</strong> the information on which these densities are based is uncertain, the distance<br />

surveyed in the Duke <strong>of</strong> Yorks, for example, is something <strong>of</strong> a semi-educated guess. <strong>The</strong><br />

definition <strong>of</strong> what is a Lapita site varies between researchers too, as does survey intensity.<br />

<strong>The</strong>re is full recognition that these figures have a high potential error and bias. That does<br />

not mean they are meaningless though, as they result from a review <strong>of</strong> the best information<br />

we have on the subject, the accumulated work <strong>of</strong> pr<strong>of</strong>essional archaeologists.<br />

<strong>The</strong>re is possibly a hint <strong>of</strong> geographic decay in the data, as the lowest useful<br />

densities were for Nissan and Buka, both to the south <strong>of</strong> the other near-Oceanic data, but<br />

this may simply be a sample-size effect given the low number <strong>of</strong> observations, or may<br />

157


Figure 6: Roviana and Kaliquongu surveys as samples <strong>of</strong> the New Georgia lagoon and<br />

barrier island system.<br />

relate to survey methodology and/or the effects <strong>of</strong> preservation and visibility on probability<br />

<strong>of</strong> detection.<br />

Sampling <strong>The</strong>ory and the Roviana/Kaliquongu Survey Regions:<br />

Orton notes that an archaeological sampling region may be <strong>of</strong> any size, but should<br />

have some geographic or cultural coherence (Orton 2000: Chapter 4). <strong>The</strong> Roviana and<br />

Kaliquongu surveys reported below can be seen as two single building block sample<br />

populations in two de facto regional sampling strategies, but what is the target population<br />

(Orton 2000:18) being sampled? <strong>The</strong>y could be viewed as samples <strong>of</strong> the Lapita/Post<br />

Lapita culture region, which stretched (according to present understandings <strong>of</strong> Lapita)<br />

from the Bismarck Archipelago to Tonga-Samoa. Roviana can also be considered as a<br />

sample <strong>of</strong> the near-Oceanic biogeographic region, in extent from the Bismark Archipelago<br />

158


to Makira/Ulawa. Perhaps the Roviana survey is best considered as a sample <strong>of</strong> the<br />

geographically, geologically and tectonically coherent lagoon circuit <strong>of</strong> New Georgia<br />

(Figure 6). Ultimately the significance <strong>of</strong> the Roviana results need to be considered at<br />

more than one <strong>of</strong> these scales in comparison with other “building block” surveys.<br />

This raises questions <strong>of</strong> comparability, and highlights the importance <strong>of</strong> the details<br />

<strong>of</strong> survey methodology and assessments <strong>of</strong> archaeological visibility and archaeological<br />

preservation for comparative data. In this respect, the Roviana survey is more comparable<br />

to most other regional surveys than the Kaliquongu survey at the broader scales <strong>of</strong><br />

comparison mentioned above. <strong>The</strong> Kaliquongu survey is in some respects comparable to<br />

the recent Buka survey (Wickler 1995) in that Wickler states that examination <strong>of</strong> reef flats<br />

was purposive, but information in that case on site visibility and quantification <strong>of</strong> survey<br />

coverage was not given, limiting comparative utility.<br />

Preservation and Visibility:<br />

In contrast to the Bismark archipelago, the geographic and geological setting <strong>of</strong> the<br />

Roviana/Kaliquongu surveys is an area <strong>of</strong> low seismicity, and despite the presence <strong>of</strong><br />

active undersea volcanism <strong>of</strong>f southern New Georgia, is not blanketed with any airfall<br />

tephras. Recent work explicating sea-level history (Mann et al. 1998) suggests sherd<br />

scatters in the intertidal zone are currently in the process <strong>of</strong> emerging from the sea into the<br />

swash zone, and being destroyed in the process. <strong>The</strong> relatively high visibility <strong>of</strong><br />

archaeological sites in the intertidal zone <strong>of</strong> coralline gravel shorelines and the ease with<br />

which sherd samples can be obtained enable a regional spatial sampling approach along<br />

such shorelines with the aim <strong>of</strong> obtaining intersite distributional information. <strong>The</strong>re are<br />

other s<strong>of</strong>t-sediment shorelines bordering these lagoons for which archaeological visibility<br />

is much reduced. One might expect preserved ceramic/lithic scatters to be obscured to<br />

159


some degree along such s<strong>of</strong>t-sediment shorelines.<br />

Even along the sheltered shorelines <strong>of</strong> the lagoon, archaeological preservation in<br />

zones <strong>of</strong> high intertidal site visibility is less than ideal, as evidenced by estimates <strong>of</strong> the<br />

parent vessel populations from which sherd samples derive (see Chapter 5), and in general<br />

the recovered samples seem to only be a tiny fraction <strong>of</strong> the absolute quantities <strong>of</strong> material<br />

in the minimum breakage population. <strong>The</strong>re is thus great potential for site density in the<br />

zone <strong>of</strong> high visibility to under-represent total site density, although some lithic lag-deposit<br />

scatter should outlast the ceramics (but these are not chronologically diagnostic to the<br />

same extent as the ceramics).<br />

Both the Roviana and Kaliquongu surveys comprise sample regions that are<br />

geographically most directly representative <strong>of</strong> the lagoon systems <strong>of</strong> the New Georgia<br />

Group. While a feature <strong>of</strong> the geology <strong>of</strong> the New Georgia group is very substantial and<br />

rapid uplift <strong>of</strong> the forearc region (Ranongga/Southern Rendova/Tetepare), the bulk <strong>of</strong> the<br />

group, and particularly the main island <strong>of</strong> New Georgia, is more stable, surrounded by an<br />

intricate system <strong>of</strong> elevated Pleistocene-age coral reefs and lagoons (Dunkley 1986, Mann<br />

et al. 1998, Stoddart 1969a, b). As discussed in Chapter 1, this system <strong>of</strong> lagoons,<br />

mainland shorelines and barrier islands forms a near-continuous circuit <strong>of</strong> New Georgia,<br />

and is relatively unvarying in terms <strong>of</strong> the diversity <strong>of</strong> preservational and visibility<br />

zonation over much <strong>of</strong> its length, although uplift is less at Marovo Lagoon. <strong>The</strong> area<br />

covered by these lagoons and elevated barrier islands and the crenulate drowned mainland<br />

shoreline total over 900km 2 (Stoddart 1969a: 387) (15,000km 2 including New Georgia<br />

mainland and the surrounding waters as measured in the review). From the presently<br />

published geological/geomorphological literature it would seem that the 200km 2 Roviana<br />

lagoon is roughly comparable as a settlement/preservation/visibility context to the 700km 2<br />

<strong>of</strong> barrier lagoons along the whole north-east coast <strong>of</strong> New Georgia, comprising Tongavae<br />

160


Lagoon, Ngerrasi Lagoon, Marovo Lagoon, Kolo Lagoon and Kalikolo Lagoon, and also<br />

the Panga Bay area south <strong>of</strong> Vangunu, and can be reasonably regarded as a 2/9 sample <strong>of</strong><br />

this area. <strong>The</strong> Kaliquongu intertidal archaeological survey is best seen as most directly<br />

representative <strong>of</strong> intertidal patterning within this 900km 2 area, and results should only be<br />

extrapolated to other environmental contexts with caution.<br />

In characterizing the archaeological distribution in a sample region, the aim is to<br />

obtain estimates <strong>of</strong> the values <strong>of</strong> variables that can be measured in principle on our target<br />

region (Orton 2000:18). For the Roviana and Kaliquongu surveys aims were to obtain<br />

estimates <strong>of</strong> the ratio variable site density (number <strong>of</strong> sites per km <strong>of</strong> coastline and number<br />

<strong>of</strong> sites per square kilometre) for two coarsely defined periods, Lapita and Post-Lapita (as<br />

measured by ceramic styles).<br />

In summary, features <strong>of</strong> this environment pertinent to the approach taken are:<br />

• reduced wave-exposure through the protection <strong>of</strong> the raised coral barrier<br />

islands, leading to enhanced preservation <strong>of</strong> intertidal sites,<br />

• a low degree <strong>of</strong> sedimentation on the barrier island shorelines, resulting in<br />

high probability <strong>of</strong> site detection in some locations given appropriate<br />

survey methods and reasonable preservation. Additionally, the use <strong>of</strong> a<br />

single sample region for each <strong>of</strong> the two survey methods means the results<br />

do not comprise estimates <strong>of</strong> New Georgia site density in a statistical sense.<br />

One further set <strong>of</strong> geological features worthy <strong>of</strong> particular note is the number and location<br />

<strong>of</strong> deep-water reef passages between the barrier islands. <strong>The</strong>re are around 40 passages in<br />

total but many are shallow, and the deeper <strong>of</strong> these may be favourable settlement<br />

localities, due to navigational advantages and marine biodiversity. Reef passages give<br />

161


access to pelagic resources in the channels surrounding New Georgia, as the outer coasts<br />

<strong>of</strong> barrier islands are (in general) unsuited to settlement (there are modern exceptions, for<br />

example Himo bay in the Saikile district) <strong>The</strong>se ge<strong>of</strong>acts and bi<strong>of</strong>acts are relevant to any<br />

archaeological sampling <strong>of</strong> the lagoon system, as their influence on late-Holocene site<br />

location is likely to be pr<strong>of</strong>ound.<br />

Data Quality:<br />

<strong>The</strong> aim <strong>of</strong> acquiring estimates <strong>of</strong> site density depends not only on good archaeological<br />

data from Roviana Lagoon, but also on a repeatable measure <strong>of</strong> survey area and distance.<br />

Shoreline length is a more objective measure than gross area, as there are a number <strong>of</strong><br />

problems in defining the areal limits <strong>of</strong> survey in an islandic environment, and the land-to-<br />

sea ratio and area-to-coastline ratio will vary from place to place in such environments.<br />

Shoreline length is regarded as a useful predictive unit for the New Georgia lagoon<br />

intertidal target region, with reef-passage count less clear (not all reef passages were<br />

created equal, despite use <strong>of</strong> this measure elsewhere (Felgate 2002).<br />

Probability <strong>of</strong> Detection is a major archaeological sampling issue and source <strong>of</strong> non-<br />

sampling errors (Orton 2000:26). Sites on surveyed reef flats or gravelly sediments were<br />

considered to have an equal probability <strong>of</strong> detection, although variation in the degree <strong>of</strong><br />

preservation could complicate chronological assignment. Sites on s<strong>of</strong>t-sediment shores<br />

were considered to have a lower and more variable probability <strong>of</strong> detection. <strong>The</strong> absolute<br />

values <strong>of</strong> these probabilities can only be guessed at, a problem by no means unique to the<br />

Roviana survey (Orton 2000:39), so that no standard errors can be calculated. For hard-<br />

sediment shorelines the probability <strong>of</strong> detection <strong>of</strong> extant deposits is probably close to<br />

100%, although if taphonomic processes have removed ceramic stylistic information the<br />

162


probability <strong>of</strong> detection by period is less than 100%, while for s<strong>of</strong>t sediments it is probably<br />

near to zero, unless as findspots.<br />

If Lapita does favour reef passage locations, which tend to have hard-sediment<br />

shorelines and high probability <strong>of</strong> detection (assuming sufficient preservation), this might<br />

lead to a higher probability <strong>of</strong> detection for this site type in comparison to more widely<br />

distributed site types.<br />

Survey Methods:<br />

Design <strong>of</strong> the Roviana and Kaliquongu surveys is considered in relation to the 12 stages<br />

<strong>of</strong> sample survey recommended by Orton (Orton 2000:27, 76-100). <strong>The</strong>se are:<br />

1. assimilation <strong>of</strong> existing knowledge<br />

2. objectives <strong>of</strong> the survey<br />

3. population to be sampled<br />

4. data to be collected<br />

5. degree <strong>of</strong> precision required<br />

6. method <strong>of</strong> measurement<br />

7. the frame<br />

8. selection <strong>of</strong> the sample<br />

9. the pre-test<br />

10. organization <strong>of</strong> the fieldwork<br />

11. summary and analysis <strong>of</strong> the data<br />

12. information gained for future surveys<br />

163


Methods <strong>of</strong> the Roviana Survey:<br />

Assimilation <strong>of</strong> existing knowledge:<br />

What do existing surveys tell us about how much work needed to get the information<br />

desired? It has been suggested throughout this <strong>chapter</strong> that pioneering informant-<br />

prospection surveys <strong>of</strong> the type conducted on Bougainville by Terrell and in Roviana<br />

Lagoon by the NGAS have a late-site bias. This suggests that a substantial amount <strong>of</strong> work<br />

would need to be done to fortuitously discover a proportion <strong>of</strong> early sites. Terrell’s 86km<br />

<strong>of</strong> Bougainville coastline surveyed in this manner failed to locate early (Lapita-era) pottery<br />

sites (either because there never were any there or because they were not<br />

preserved/detected by the scale and method <strong>of</strong> survey), as did Irwins’ Shortland survey,<br />

although in the Shortlands case it must be noted that Irwin targeted terrestrial reef passage<br />

locations, which may have improved the probability <strong>of</strong> early site detection. Current<br />

evidence from Buka and Roviana suggests that a) this method will have a low return as a<br />

prospection technique for Lapita sites and b) even large surveys such as Bougainville may<br />

yield no return for the early ceramic period.<br />

<strong>The</strong> Roviana Lagoon has an inner coastline length <strong>of</strong> 180km (estimated by walking<br />

a 1km divider along the mapped coastline at 1:50 000 scale) and therefore is double the<br />

amount <strong>of</strong> coastal survey conducted by Terrell by similar methods on Bougainville.<br />

Additionally, this lagoon shoreline is not exposed to tsunami or cyclone storm swells<br />

(discussed in detail in Chapter 6), although storm surge associated with cyclones passing<br />

to the south through the Bellona area does affect water-levels in the lagoon, flooding the<br />

lower shore platforms on occasion (personal observation, 1998).<br />

164


<strong>The</strong> Roviana survey therefore represents a significant addition to the knowledge-set gained<br />

through informant-prospection and fortuitous observation coastal survey methods.<br />

Objectives <strong>of</strong> the Roviana survey:<br />

<strong>The</strong> stated objectives <strong>of</strong> the Roviana Archaeological survey were to conduct a preliminary<br />

archaeological survey <strong>of</strong> the region to determine the location and type <strong>of</strong> sites present in<br />

this region (Sheppard 1996:4). Clearly these are not density estimation objectives (Orton<br />

2000:79). For the purposes <strong>of</strong> the present thesis these objectives are recast as :<br />

to test the hypothesis <strong>of</strong> Lapita settlement <strong>of</strong> New Georgia in the past, and to gain<br />

if possible an estimate <strong>of</strong> Lapita site density. Further, to estimate the density <strong>of</strong> other early<br />

post-Lapita sites known to be located across the target region.<br />

Population to be sampled by the Roviana archaeological survey:<br />

<strong>The</strong> hypothetical Lapita sites <strong>of</strong> the coastlines <strong>of</strong> the New Georgia lagoon system was the<br />

target population to be sampled. This region comprises an area <strong>of</strong> 900km 2 approximately<br />

(Stoddart 1969a), and is geologically and geographically homogenous and bounded<br />

through relative uniqueness in these respects in the broader region <strong>of</strong> near-Oceania. This<br />

was a surface survey, and for the present purposes, a survey <strong>of</strong> early ceramic sites (the<br />

majority <strong>of</strong> archaeological evidence resulting from the 1996 field seasons was <strong>of</strong> other<br />

types <strong>of</strong> site, but these do not directly relate to the present objectives). This region was<br />

considered as a series <strong>of</strong> environmental zones between which site obtrusiveness,<br />

preservation and visibility could be expected to vary. <strong>The</strong>se were:<br />

165


• the mainland shoreline, consisting for the most part recent alluvium<br />

• the mainland intertidal, consisting for the most part s<strong>of</strong>t sediments with high<br />

preservation potential, low archaeological visibility for the sought-after site types,<br />

and low site obtrusiveness;<br />

• the intertidal zone <strong>of</strong> barrier-island shorelines, predominantly although with major<br />

exceptions sandy to coralline gravel/ coralline rocky shorelines, with low site<br />

obtrusiveness, good archaeological visibility and moderate preservation; the<br />

terrestrial coastlines <strong>of</strong> the barrier islands, the recent emerged coral/reefal detritus<br />

shore platform with for the most part fairly dense vegetation, low site visibility and<br />

moderate preservation potential due to modern and possibly prehistoric gardening<br />

activities.<br />

Data collected:<br />

<strong>The</strong> primary unit <strong>of</strong> measurement was the site, with <strong>of</strong>fsite data recorded as findspots. For<br />

site I use Schiffer’s definition: “a high density area <strong>of</strong> artefacts (Schiffer et al. 1978:2,14).<br />

By findspot I mean the location <strong>of</strong> a group <strong>of</strong> three or less sherds in an area <strong>of</strong> otherwise<br />

low ceramic density. Aceramic lithic scatters are similarly regarded as sites, while aceramic<br />

findspots were not recorded. Late-Lapita sites were initially defined as those which yielded<br />

ceramics <strong>of</strong> multi-part slab vessel construction and Lapita design structure (defined in<br />

detail in Chapters 4, 8,9 and 12) (see Figure 7, Figure 8, Figure 9).<br />

166


Figure 7: Slab-built carinated vessels from Honiavasa initially assigned to the Lapita<br />

period (this assignment is examined in more detail in Chapters 8 and 9): the solid<br />

vertical line in HV.2.464 represents an estimated location <strong>of</strong> the central vertical avis<br />

(CVA) <strong>of</strong> the pot.<br />

167


Figure 8: Carinated vessels from Honiavasa, showing double-line markes on some<br />

design zones, bands <strong>of</strong> nubbins at the neck in two cases, and a band <strong>of</strong> fingernail<br />

impression in one case (top).<br />

168


Figure 9: HV.4.202 has an incised motif laid out in double lines; HV.1.314 is dentatestamped,<br />

with similar design structure and a double carination; HV.2.341 is carinated,<br />

with the design laid out in applied fillets, bounded horizontally by decorated lap joins<br />

between slabs; HV.2.297 and HV.4.379 have bands <strong>of</strong> single fingernail impressions and<br />

nubbins at the neck respectively.<br />

169


Sites with mainly simple one-piece globular forms, together with Miho-style or<br />

Gharanga/Kopo-style decoration and rim forms (decorative styles named for sites in which<br />

they were first collected in quantity, locations shown below) were classified for the coarse<br />

periodization <strong>of</strong> the current <strong>chapter</strong> as post-Lapita. This classification is examined in detail<br />

in other <strong>chapter</strong>s. <strong>The</strong>se styles are defined below as follows:<br />

• Miho style includes a diversity <strong>of</strong> vessel forms, but sharply carinated forms are<br />

virtually absent, and simple one-piece pots generally have a thick neck decorated<br />

most commonly with a band <strong>of</strong> opposed-pinch fingernail impression. Decorative<br />

techniques include also incision, perforation and applique, and incised motifs are<br />

usually made up <strong>of</strong> triangular secondary design zonation within a primary design<br />

zone band, constructed with single lines rather than double. <strong>The</strong>re is no incised<br />

horizontal upper or lower boundary to the primary design zone band. On the rim<br />

the triangles are usually filled by parallel lines, but are unfilled on the shoulder.<br />

Examples are illustrated in Figure 10, Figure 11, Figure 12, Figure 13 and<br />

Figure 14.<br />

• Gharanga/Kopo style can be defined for the present purposes by three diagnostic<br />

attributes, a band <strong>of</strong> punctation in the neck region and/or multiple bands <strong>of</strong><br />

opposed-pinch fingernail impression on the shoulder, and/or a short, heavily<br />

everted rim. Examples are illustrated in Figure 15, Figure 16, Figure 17, Figure<br />

18, Figure 19, Figure 20, Figure 21, Figure 22, Figure 23, Figure 24, Figure<br />

25, and Figure 26.<br />

170


Figure 10: Incised rims with opposed-pinch fingernail impressed band at the neck,<br />

diagnostic <strong>of</strong> the Miho subgroup <strong>of</strong> Post-Lapita styles.<br />

171


Figure 11: Miho-style post-Lapita sherds.<br />

Figure 12: Miho-style post-Lapita sherds with incised shoulders.<br />

172


Figure 13: Miho-style post-Lapita sherds with CVA measurements shown (MH290<br />

was measured at to locations on the pr<strong>of</strong>ile; at the interior <strong>of</strong> the neck orifice and the<br />

exterior shoulder).<br />

173


Figure 14: Miho-style post-Lapita sherds; dashed CVA lines are measurements based<br />

on non-circular (uneven) curvature at the pr<strong>of</strong>ile locations indicated by arrows.<br />

174


Figure 15: Gharanga-style post-Lapita sherds. (Gharanga is a short-rim subgroup<br />

<strong>of</strong> Gharanga-Kopo which may have multiple bands <strong>of</strong> opposed-pinch fingernail<br />

impression on the shoulder.<br />

Figure 16: Gharanga-style post-Lapita sherds.<br />

175


Figure 17: Gharanga/Kopo style post-Lapita sherds (top) and a large Gharanga-style<br />

sherd (bottom) (four measurement points used as arrowed to estimate CVA.<br />

176


Figure 18: Intermediate between Gharanga and Kopo styles: all post-Lapita.<br />

Figure 19: Kopo-style post-Lapita rim sherds.<br />

177


Figure 20: Another large Gharanga-style post-Lapita sherd from an unrestricted vessel<br />

form.<br />

Figure 21: Gharanga-style small-orifice post-Lapita sherd showing rolled rim and thin<br />

wall common to this style.<br />

Figure 22: Kopo-Style post-Lapita sherd (taller, less everted rim than Gharanga style)<br />

from a large-orifice vessel, with bands <strong>of</strong> impressions along both inner and outer edges<br />

<strong>of</strong> the lip.<br />

178


Figure 23: Less decorated variant <strong>of</strong> Gharanga-Kopo post-Lapita style, without a<br />

strong corner point in vertical section at the neck.<br />

Figure 24: Large-orifice Gharanga/Kopo-style sherd with deformation <strong>of</strong> the lip into a<br />

wave pattern.<br />

Figure 25: Large Gharanga-style post-Lapita sherd with typical decoration, including a<br />

band <strong>of</strong> impressions along the inner edge <strong>of</strong> the lip.<br />

179


Figure 26: Gharanga/Kopo-style post-Lapita vessels: most are weakly restricted at the<br />

neck, with short, heavily everted rims. Punctate band at the neck is the most common<br />

decoration in this group, while multiple bands <strong>of</strong> fingernail pinch are common on the<br />

short-rim examples. One sherd (MH.33) had exotic quartz-calcite hybrid temper (see<br />

Chapter 4).<br />

180


Most collection sites yielded a mixture <strong>of</strong> these two periods to a slight extent (this will be<br />

discussed in detail in Chapter 12), but showed predominance <strong>of</strong> one period in the styles<br />

represented.<br />

Degree <strong>of</strong> precision required:<br />

Precision could not be calculated due to use <strong>of</strong> a single sample region.<br />

Method <strong>of</strong> measurement:<br />

Analysis <strong>of</strong> ceramic variability <strong>of</strong> surface collections was used to assign sites to period. Site<br />

locations were recorded on 1:50000 topographic maps. Site visibility for the intertidal zone<br />

was classified as high or low using 1:25000 air photographs (gravelly reef flats and coral<br />

sand flats show pale or white while s<strong>of</strong>t sediments show dark) combined with ground<br />

truthing. Site density was calculated for both the total 180km <strong>of</strong> shoreline and for the<br />

70km <strong>of</strong> high-visibility shoreline<br />

<strong>The</strong> frame:<br />

<strong>The</strong> sampling frame for the Roviana survey was as shown in Figure 6.<br />

Selection <strong>of</strong> the sample:<br />

Survey was conducted on land and in the sea as directed by informants.<br />

<strong>The</strong> pre-test:<br />

No formal pre-test was conducted, other than the informant-prospecting survey reported<br />

by Reeve (Reeve 1989)<br />

Organization <strong>of</strong> the fieldwork:<br />

Sheppard, Felgate, Keopo and Roga spent five weeks in the Kaliquongu area <strong>of</strong> the<br />

Lagoon in January-February <strong>of</strong> 1996, using the informant-prospection method to locate<br />

traditional sites, rockshelters and early ceramic sites and findspots. A further season <strong>of</strong><br />

similar fieldwork was spent in the Saikile area, including participation by Sheppard,<br />

Walter, Felgate, Roga, Jones, Nagaoka and informants, most notably Mr.Sae Oka <strong>of</strong><br />

181


Patmos, Nusahope area. An observer-bias in the results is likely in that Mr.Sae Oka had<br />

developed a particular interest in intertidal pottery scatters and actively located a number<br />

<strong>of</strong> these on the Ndora Island coastline, while in the Kaliquongu area the informants initially<br />

consulted reported only one such locality, at the mouth <strong>of</strong> a mainland stream, Gharanga,<br />

where a landowner (Mr Phillip Lanni) had developed a personal interest in the pottery and<br />

artefacts noticed during gardening/copra activities at that specific locality (the stream is<br />

also used today to obtain water in the dry season).<br />

Summary and analysis <strong>of</strong> the data:<br />

Overall survey results supported the pattern noted previously (Reeve 1989) <strong>of</strong> early<br />

ceramic evidence located exclusively in the intertidal zone. Thin plain ceramics were found<br />

on land, associated with late-prehistoric ritual stone structures. Similar sherds were found<br />

in an inland midden deposit on the mainland dated to 1403-1490AD at one sigma (Site 25<br />

in Sheppard 1996). Only sites with sufficient ceramic evidence to allow coarse temporal<br />

characterization were counted in the analysis. Findspots comprised sites with three or less<br />

sherds in the general locality. No aceramic lithic scatters were noticed during the survey.<br />

Lapita sites were not initially located, although in a subsequent field season the Nusa<br />

Roviana site was located, which had one sherd with dentate-stamping. Results were<br />

initially interpreted (prior to the Kaliquongu survey and the Nusa Roviana dentate find) as<br />

suggesting Lapita was absent in the New Georgia group, and possibly elsewhere in the<br />

near-Oceanic Solomon Islands (Sheppard et al. 1999). A number <strong>of</strong> post-lapita sites were<br />

located, including Hoghoi, Paniavile (previously recorded by Reeve), Ririgomana, Punala,<br />

Humbi Quongu, Rangu, Gharanga, and Kopo. Intertidal ceramic findspots were noted at<br />

Mare point, Kazu, Pikoro, Mbaraulu, and Sasavele (Figure 27).<br />

Information gained for future surveys:<br />

Early ceramics seemed exclusively intertidal, while late ceramics were principally found<br />

associated with terrestrial ritual structures although a broader scatter <strong>of</strong> tiny fragments <strong>of</strong><br />

this thin pottery was noted in garden areas.<br />

182


Munda<br />

Malangari<br />

Island<br />

183<br />

Saikile<br />

Figure 27: Locations mentioned in the text in relation to Roviana and<br />

Kaliquongu surveys.<br />

<strong>The</strong> Kaliquongu Survey:<br />

Objectives <strong>of</strong> the survey:<br />

A primary reason for surveying in 1997-8 in the Kaliquongu region <strong>of</strong> Roviana Lagoon<br />

was to test the hypothesis that Lapita pottery sites were absent. Additional objectives were<br />

to describe the distributions <strong>of</strong> archaeological materials by stylistic period. An important<br />

secondary objective was methodological, and in this regard the Roviana survey is regarded<br />

as an experiment in intertidal-zone/shallow-water survey and sampling. What would the<br />

results <strong>of</strong> such a survey be? What implications would these have for archaeology <strong>of</strong> the<br />

Lapita era in Near-Oceania?


Population to be sampled:<br />

As for the Roviana survey, the Kaliquongu survey region was regarded as geographically<br />

representative <strong>of</strong> the New Georgia lagoonal system. <strong>The</strong> populations to be sampled were<br />

the intertidal/shallow water Lapita sites <strong>of</strong> New Georgia, and the intertidal post-Lapita<br />

early pottery sites.<br />

Data to be collected:<br />

<strong>The</strong> primary unit <strong>of</strong> measurement was the site, with <strong>of</strong>fsite data recorded as findspots. Sites<br />

were defined as high density scatters <strong>of</strong> artefacts, sufficiently numerous to allow a coarse<br />

ceramic periodization. This decision was taken with recording technology in mind: the<br />

relative remoteness <strong>of</strong> the location and lack <strong>of</strong> availability <strong>of</strong> electronic recording aids or<br />

power supplies dictated a reduction in the spatial resolution <strong>of</strong> data recording below what<br />

is fast becoming the norm in distributional or <strong>of</strong>f-site archaeology, i.e. point-provenancing<br />

<strong>of</strong> artefacts. Artefacts were collected rather than documented in situ, both to prevent rare<br />

categories <strong>of</strong> finds being missed due to marine encrustation, and to enable development <strong>of</strong><br />

classificatory systematics in the laboratory. Some sites were collected as single samples,<br />

some as samples from unequal sub-areas, one as a linear transect <strong>of</strong> 5x12m “squares”, and<br />

one as a series <strong>of</strong> 1m transects at 5m intervals, each comprising a linear arrangement <strong>of</strong><br />

5x1m “squares”.<br />

Degree <strong>of</strong> precision required:<br />

In terms <strong>of</strong> the hypothesis that Lapita is/is not present in the Western Solomon Islands, the<br />

184


prime difficulty in specifying the degree <strong>of</strong> precision required arises from unknown<br />

probability <strong>of</strong> detection. <strong>The</strong> existing knowledge used to formulate the expected Lapita site<br />

density was not sufficiently detailed regarding factors contributing to site detectability to<br />

allow any precise estimate <strong>of</strong> the size <strong>of</strong> a survey region required to establish the absence<br />

<strong>of</strong> Lapita. In this regard bigger is better theoretically, but in practice resources were<br />

limited, as was access, imposing limits on the size <strong>of</strong> the sample region that could be<br />

accessed and surveyed intensively in a manner that allowed biases to be assessed. <strong>The</strong><br />

review <strong>of</strong> 17 coastal surveys given above suggests for high-intensity survey Lapita site<br />

density could be around 91 sites per 1000km <strong>of</strong> coastline or upwards, or on average at<br />

least one site per 11km <strong>of</strong> coastline. This suggests that the survey region would need to<br />

be greater than 11km <strong>of</strong> high-detectability coastline favourable to Lapita settlement to<br />

have a reasonable chance <strong>of</strong> locating a Lapita site at the minimum recorded site density for<br />

intensive survey.<br />

Method <strong>of</strong> Measurement:<br />

Site locations were recorded on air photographs or 1:50 000 topographic maps, augmented<br />

on featureless coastlines with compass bearings to prominent landmarks, usually to nearby<br />

islets in the lagoon. Artefacts/sherds and manuports were bagged by intrasite transect or<br />

in the absence <strong>of</strong> transect collection, by site. Transects were set out using fibreglass<br />

measuring tapes and marked out with nylon mon<strong>of</strong>ilament on stakes. Field desalination and<br />

laboratory cleaning were important aspects <strong>of</strong> measurement, particularly <strong>of</strong> sherd<br />

decoration and temper characterization. Site density was calculated in relation to the<br />

185


distance <strong>of</strong> high-visibility shoreline surveyed, and as number <strong>of</strong> sites/findspots/lithic scatters<br />

per square kilometre.<br />

<strong>The</strong> Frame:<br />

<strong>The</strong> frame was as shown in Figure 6, which comprised the intertidal zone <strong>of</strong> the<br />

Kaliquongu region <strong>of</strong> the lagoon, chosen for reasons <strong>of</strong> access and because this appeared<br />

to be a favourable environment for settlement due to presence <strong>of</strong> a deepwater barrier-island<br />

passage and harbour, many small islets, reef shallows and seagrass flats, and a variety <strong>of</strong><br />

barrier-island and mainland resources within easy paddling distance <strong>of</strong> each other (Figure<br />

28). This area encompassed 70km 2 , approximately one third <strong>of</strong> Roviana Lagoon and<br />

approximately 8% <strong>of</strong> the total 900km 2 <strong>of</strong> the New Georgia lagoonal system.<br />

Selection <strong>of</strong> the sample:<br />

<strong>The</strong> sampling transect comprised the firm-sediment walkable sections <strong>of</strong> a natural<br />

landscape unit, the intertidal zone, as shown on Figure 28. <strong>The</strong> sampling fraction was<br />

intended to be 100% <strong>of</strong> the intertidal zone within the survey limits, but in practice some<br />

areas were impassable, and others were not covered in the time available. About 15km <strong>of</strong><br />

high-visibility intertidal was surveyed in this manner, and a further 10km or so <strong>of</strong><br />

mangroves and s<strong>of</strong>t sediments was surveyed where passable.<br />

<strong>The</strong> Pre-test:<br />

<strong>The</strong> 1996 field seasons had effectively provided a pre-test <strong>of</strong> methods. Site location<br />

186


technique was initially predominantly through reports by interested local residents or<br />

through fortuitous discovery during the process <strong>of</strong> traveling to villages by canoe to consult<br />

with communities regarding the archaeological survey. This was regarded as<br />

unsatisfactory, and the determination to take a more systematic sampling approach to the<br />

landscape developed out <strong>of</strong> it. <strong>The</strong> January/August surveys <strong>of</strong> 1996 involved a substantial<br />

terrestrial component, as did later work, which had shown a dominant terrestrial pattern<br />

<strong>of</strong> a thin, largely plain pottery style associated with late-prehistoric archaeological evidence<br />

(Sheppard et al. 1999:319).<br />

Organization <strong>of</strong> the Fieldwork:<br />

Survey was conducted during the season <strong>of</strong> low tides at the low-water spring tides <strong>of</strong> each<br />

month. This had a major beneficial effect on site detectability. Sites were surface-collected<br />

in most cases for laboratory classification <strong>of</strong> site type. Fieldwork assistants and guides<br />

were recruited from Sasavele village for the Kaliquongu survey and surface collections.<br />

Summary and Analysis <strong>of</strong> the Data:<br />

This was as for the Roviana survey, except site visibility was noted qualitatively in the field<br />

during survey rather than assessed from air photographs.<br />

Information Gained for Future Surveys:<br />

Time and cost information, intertidal site detectability under different conditions, intrasite<br />

archaeological potential and the effects on this <strong>of</strong> grab sampling, and the solutions to<br />

187


technical and logistical problems encountered during the survey, should all be <strong>of</strong> interest<br />

to archaeologists interested in the Lapita ceramic series in near-Oceania. This information<br />

will be <strong>of</strong> use to future researchers and is regarded as <strong>of</strong> immediate importance to cultural<br />

resource managers in the region.<br />

Results:<br />

Results <strong>of</strong> the Kaliquongu survey are shown in Figure 28. In tabulating these results<br />

(Table 6), two site types were lumped, aceramic lithic manuport scatters, and lithic<br />

scatters with rare ceramic finds. Two <strong>of</strong> the aceramic scatters recorded during the<br />

Kaliquongu survey may be <strong>of</strong> fluvial rather than anthropogenic origin, so the site density<br />

<strong>of</strong> these site types may be too high.<br />

Table 6: Site densities by period for Roviana and Kaliquongu surveys.<br />

Site Type Roviana<br />

High vis.<br />

(70km)<br />

Roviana<br />

low vis.<br />

(110km)<br />

Roviana<br />

Total<br />

(180km)<br />

188<br />

Kaliqu.<br />

High vis.<br />

(15km)<br />

Lapita 0 0 0 67/1000km<br />

n=1<br />

Post-<br />

Lapita<br />

129/1000km<br />

n=9<br />

9/1000km<br />

n=1<br />

56/1000km<br />

n=10<br />

333/1000km<br />

n=5<br />

aceramic 0/1000km 0 0 267-<br />

400/1000km<br />

n=4-6<br />

findspots 43/1000km<br />

n=3<br />

0 17/1000km<br />

n=3<br />

Kaliqu.<br />

Low vis.<br />

(10km)<br />

0 200/1000km<br />

n=2<br />

Kaliqu<br />

Total<br />

(25km)<br />

0/1000km 40/1000km<br />

n=1<br />

0/1000km 200/1000km<br />

n=5<br />

0 240/1000km<br />

n=6<br />

80/1000km<br />

n=2


Figure 28: Kaliquongu survey transects: the unfilled symbols represent sites discovered<br />

by informant-prospection during the Roviana survey.<br />

189


Discussion and Conclusions: the Two Surveys.<br />

Disregarding the emphasis on intertidal ceramics, and considering the Lapita site density<br />

figures obtained, the results <strong>of</strong> the Roviana survey do not seem significantly different to<br />

those <strong>of</strong> Terrell’s Bougainville survey. Where Terrell surveyed 86 km <strong>of</strong> coastline by the<br />

informant prospection method and found no Lapita, the Roviana Survey covered 180km<br />

<strong>of</strong> Roviana lagoon coastline by similar methods and found a site with a single dentate sherd<br />

(Nusa-Roviana-Zoraka). Neither do the results seem all that significantly different to<br />

results from Bellona, Guadalcanal or Santa Cruz. It must be borne in mind that the results<br />

<strong>of</strong> extensive, low-intensity survey using informant prospection will be highly variable<br />

depending on informant selection, as evidenced by clustered site distribution in the Roviana<br />

survey results (Sae Oka’s site discoveries on Ndora Island for example). Also, factors<br />

such as coastal site preservation and particularly intertidal site preservation, have not<br />

entered in any systematic way into discussions to date. Locating a single Lapita site in a<br />

survey is awkward in that site density calculation clearly suffers from inadequate site<br />

sample size. <strong>The</strong> distances surveyed in the Kaliquongu survey are low, resulting in this<br />

potential source <strong>of</strong> sampling error.<br />

Taking these potential sources <strong>of</strong> error in the Roviana and Kaliquongu survey data<br />

into account, and bearing in mind also the imprecise abstraction <strong>of</strong> the comparative data<br />

from the survey results <strong>of</strong> others, it seems impossible to conclude there is evidence for any<br />

significant difference in terms <strong>of</strong> site density between the Kaliquongu survey and the<br />

general Lapita record elsewhere in Near-Oceania, particularly if one allows for an<br />

190


unknown level <strong>of</strong> site preservation. If, for example, the aceramic lithic scatters at Mbolave<br />

and Elelo were at one time Lapita sites, Kaliquongu Lapita site density would jump to 200<br />

sites per 1000km We would have to have much more extensive data from the New<br />

Georgia Lagoon system encompassing a number <strong>of</strong> sites identifiably within the Lapita<br />

series and much more intensively and systematically surveyed data from other locations in<br />

near-Oceania to demonstrate significant differences in Lapita site density.<br />

<strong>The</strong> implication for the question <strong>of</strong> the presence <strong>of</strong> other Lapita-era sites in the<br />

New Georgia group is clear. A larger survey frame, with careful attention to models <strong>of</strong> site<br />

preservation and visibility, i.e. probability <strong>of</strong> site detection, is likely to locate additional<br />

sites from this general period. What should the sampling fraction be to obtain a sample<br />

suitable for seriation analysis? To answer this we must first ask how many Lapita sites<br />

might be expected from the New Georgia lagoon circuit in total (sampling fraction 100%)?<br />

<strong>The</strong> Kaliquongu survey <strong>of</strong> 70km 2 located one Lapita site, so a best guess for the 900km 2<br />

<strong>of</strong> coastlines and lagoons <strong>of</strong> the New Georgia lagoon circuit on current data would be<br />

around 13 sites using this measure. Similarly, roughly 65 post-Lapita sites might be<br />

expected, and about 78 aceramic sites. A corpus <strong>of</strong> ceramic samples <strong>of</strong> this magnitude<br />

would allow a much better seriation analysis, which together with testing by direct dating<br />

<strong>of</strong> styles could potentially enable a fine-grained chronology <strong>of</strong> the entire period <strong>of</strong> intertidal<br />

site formation. A chronology <strong>of</strong> this sort would allow abandonment <strong>of</strong> the sort <strong>of</strong> coarse<br />

Lapita/Post-Lapita chronological construct currently in use.<br />

It is hoped that this consideration <strong>of</strong> a sampling approach to regional<br />

archaeological questions in the near-Oceanic setting highlights the a-regional and overly<br />

191


culture-historical status quo in Lapita studies (in the published literature at least) which<br />

threatens to stifle the archaeological potential <strong>of</strong> the region by largely dismissing the bulk<br />

<strong>of</strong> the Lapita record, surface or turbated deposits, as “disturbed” and therefore <strong>of</strong> low<br />

archaeological value. It is hoped that this analysis <strong>of</strong> the Roviana and Kaliquongu surveys<br />

will promote re-evaluation <strong>of</strong> the New Georgia early-ceramic surface record as the most<br />

complete witness <strong>of</strong> the history <strong>of</strong> early ceramic variability. Such a re-evaluation <strong>of</strong> survey<br />

objectives beyond a search for “good” (stratified) sites demands as a prerequisite<br />

acceptance <strong>of</strong> the tenet that survey design matters, a simple point, but one that has largely<br />

failed to penetrate Lapita studies.<br />

Uncontrolled direct reading <strong>of</strong> site densities in a comparative manner (as done by<br />

Anderson 2002) is counterproductive when the fundamental inferential issues <strong>of</strong> site<br />

preservation and visibility, or site detectability in relation to survey methods, are simply<br />

ignored in the rush to calculate higher-level behavioural variables like migration rates. <strong>The</strong><br />

poor type and extent <strong>of</strong> our sampling becomes evident when a sample-surveying<br />

perspective is engaged. If we wish to talk about migration rates, or demographic history,<br />

or colonization rates, or the distribution <strong>of</strong> “Lapita”, or when “Lapita” ends, research<br />

design which obtains data matched to the research question is needed, which for any <strong>of</strong><br />

these behavioural questions requires that we concentrate on the difficult formational<br />

questions <strong>of</strong> evidence, such as, “what is the probability <strong>of</strong> a pottery sample being<br />

preserved in a particular location and observed by the survey method?’, rather than easy<br />

questions, like “did we find anything?<br />

192


Introduction:<br />

CHAPTER 4:<br />

CERAMICS: UNITS OF DESCRIPTION<br />

<strong>The</strong> units <strong>of</strong> description employed for ceramic form and decoration were designed to<br />

enable a detailed description <strong>of</strong> ceramic variation at the level <strong>of</strong> the sherd. <strong>The</strong> problem <strong>of</strong><br />

assessing sample formation processes required careful attention to the description <strong>of</strong> sherds<br />

as sedimentary particles. Thus sherd area, mass, thickness and composition were recorded<br />

for all sherds. Description <strong>of</strong> decoration was detailed to the level <strong>of</strong> the layout, size and<br />

spacing <strong>of</strong> decorative elements, in an attempt to capture variation to the level <strong>of</strong> individual<br />

potters if necessary, in the process <strong>of</strong> deciding on units <strong>of</strong> classification (classification is<br />

covered in Chapters 8 and 9).<br />

<strong>The</strong>re is a descriptive and classificatory divide between Lapita and post-Lapita<br />

pottery (discussed in Chapter 1), in that the Frost-Irwin attribute-combination grouping<br />

approach has generally been applied to materials post-dating Lapita, while the Mead<br />

system <strong>of</strong> classification, or its non-classificatory splitting variant, the Anson system <strong>of</strong><br />

description, have dominated Lapita studies. Studying the transition between Lapita and<br />

post-Lapita requires a descriptive system that is capable <strong>of</strong> dealing both with complex<br />

motifs and with simple geometric pattern such as bands <strong>of</strong> repeated elements.<br />

<strong>The</strong> difference between Lapita and post-Lapita is already well-understood: Lapita<br />

has complex decoration including dentate-stamping in a range <strong>of</strong> designs and a diversity<br />

<strong>of</strong> (<strong>of</strong>ten slab constructed) vessel forms, some <strong>of</strong> which are characteristic. Post-Lapita<br />

lacks these distinguishing criteria (as might pre-Lapita). If at the end <strong>of</strong> this thesis I<br />

conclude there was a transition from dentate to incised and simplification in vessel form<br />

I would be merely restating the definition <strong>of</strong> Lapita rather than attempting to understand<br />

the details <strong>of</strong> ceramic change. Description in a context <strong>of</strong> reduced decorative complexity<br />

still needs to be sensitive to the temporal dimension <strong>of</strong> variability. For Roviana pottery<br />

this required that careful attention be paid to decorative location, as the sensitivity <strong>of</strong><br />

193


simple decorative techniques and simple patterns in themselves to change was not expected<br />

to be sufficient to capture fine temporal patterning. <strong>The</strong> structure <strong>of</strong> decoration across the<br />

vessel and in relation to vessel form was therefore <strong>of</strong> interest.<br />

Understanding why change, as expressed in the historical events <strong>of</strong> pot-making,<br />

happened, requires a materialist approach to the analysis <strong>of</strong> variability (Dunnell 1986:153).<br />

An inventory <strong>of</strong> designs as in the motifs <strong>of</strong> the Anson system is not sufficient when the goal<br />

is to understand design change. <strong>The</strong> units <strong>of</strong> description must reconstruct design into<br />

component attributes that are salient to design evolution. <strong>The</strong> process <strong>of</strong> how one design<br />

may have changed into another is <strong>of</strong> interest. Also, we have no reason to expect all<br />

attributes <strong>of</strong> designs to change at the same rate. Some aspect <strong>of</strong> vessel design and<br />

decoration might be highly constrained by functional adaptation or may be changing rapidly<br />

as a result <strong>of</strong> some selective pressure. Also some form or decorative attributes may be free<br />

to change stochastically, i.e. drift, in the sense <strong>of</strong> genetic drift (Dunnell 1978, 2001).<br />

Sufficient sherds in the overall sample were large enough to permit such a<br />

structural approach to the patterned layout <strong>of</strong> elements and the structure <strong>of</strong> this patterning<br />

across the vessel, and some designs were compact and simple enough to allow an<br />

element/motif design analysis. This is not to imply that elements and motifs are always<br />

hierarchically structured: some patterns are constructed as a series <strong>of</strong> repeats <strong>of</strong> an<br />

element, while some designs are more complex, and are just inventoried, being classified<br />

in the next <strong>chapter</strong> according to criteria such as zone markers (Mead 1975:24). Vessel<br />

morphology is an aspect <strong>of</strong> design that, like decoration, may be constrained or directed in<br />

some ways but free to shift or drift in others. Analysis <strong>of</strong> vessel form, using mainly interval<br />

measures like curve template fitting, was designed to capture slight variation, and enable<br />

analysis within the theoretical framework <strong>of</strong> form variability in Chapter 1.<br />

Quantitative analysis <strong>of</strong> vessel design from sherds <strong>of</strong> broken pottery requires<br />

specification <strong>of</strong> a relationship between the pieces and inferred wholes. Some <strong>of</strong> the units<br />

<strong>of</strong> description pertain to this requirement, and are used extensively in Chapter 5 to<br />

ascertain the most appropriate method <strong>of</strong> quantification. In addition to behavioural-<br />

historical and behavioural-explanatory objectives, units <strong>of</strong> description were designed with<br />

194


the objective <strong>of</strong> understanding other non-cultural aspects <strong>of</strong> site/assemblage formation<br />

processes. Accordingly, several measures <strong>of</strong> sherd size and thickness are recorded, in<br />

more detail than usual, for all sherds, allowing detailed taphonomic study in Chapters 5,<br />

6, 7 and 11.<br />

Database Structure:<br />

Figure 29: Data structure for each sherd record; each sherd can have many<br />

records in the detail tables pertaining to the various parts <strong>of</strong> the vessel<br />

represented.<br />

<strong>The</strong> core ceramic relational database used for description <strong>of</strong> sherd attributes consisted <strong>of</strong><br />

a linked arrangement <strong>of</strong> tables, with the master record linked in a one-to-many relationship<br />

to thematic tables about the sherd (Figure 29).<br />

A master table contained data pertaining to the entire sherd. Fields <strong>of</strong> data in the<br />

master record for each sherd were about sherd identity and provenance, area, weight, date<br />

recorded, time recorded, general purpose notes, fabric description code, and Estimated<br />

Vessel Equivalent (EVE) (after Orton 1993), (this is the same measurement as Egl<strong>of</strong>f’s<br />

“Percentage Factor”) (Egl<strong>of</strong>f 1973) (Table 7).<br />

195


Table 7: Structure <strong>of</strong> master record for each sherd.<br />

FIELD PURPOSE ATTRIBUTE /<br />

UNIT<br />

sherd ID identify site, unit<br />

and sherd<br />

196<br />

7-character<br />

alphanumeric<br />

METHOD OF<br />

MEASUREMENT<br />

re-join any postrecovery<br />

sherd<br />

breaks<br />

sherd area measure sherd size cm 2 grid transparency<br />

sherd weight measure sherd size g balance<br />

date recorded quality control date automatic input<br />

time recorded quality control time automatic input<br />

notes general purpose text<br />

fabric description preliminary<br />

variability<br />

special code iterative qualitative<br />

EVE measure sherd size percent rim percent chart<br />

Vessel Family link to other sherds<br />

potentially from<br />

the same vessel<br />

vessel family<br />

identity number<br />

iterative<br />

comparison<br />

(methods detailed<br />

in Chapter 5)<br />

Data pertaining to the various vessel parts represented on the sherd was distributed<br />

across a number <strong>of</strong> tables. A table <strong>of</strong> sherd thickness for each <strong>of</strong> the vessel parts<br />

represented on the sherd was related to the master table via the sherd identity code.<br />

Similarly, a range <strong>of</strong> data about vessel form for each part <strong>of</strong> the vessel represented on the<br />

sherd was related to the master table via the sherd identity code. A fourth table comprised<br />

information about decoration on each part <strong>of</strong> the vessel, again related to the master record<br />

for the sherd via the sherd identity code. <strong>The</strong>se tables are appended on CD (Master.db,<br />

Thickness.db, Form.db and Decoration.db).<br />

Master Record Table: Explanation <strong>of</strong> Data Codes for Each Field:<br />

<strong>The</strong> sherd ID field was a 7-character assignment with the following site prefixes used:<br />

R37 Paniavile (2 nd collection) (R37 had two redundant unit characters <strong>of</strong> theform1a, 2a,


P Paniavile (1 st collection)<br />

C Paniavile (villager collection stored in cave)<br />

A Zangana (1 st collection)<br />

Z Zangana (2 nd Collection)<br />

B Zangana (3 rd collection)<br />

GH Gharanga (1 st collection)<br />

GE Gharanga (2 nd collection- east <strong>of</strong> stream)<br />

GW Gharanga (2 nd collection- west <strong>of</strong> stream)<br />

MH Miho collection<br />

HV Honiavasa collection<br />

HG99 Hoghoi (1 st collection)<br />

HG Hoghoi (2 nd collection)<br />

K Kopo collection<br />

NR Nusa Roviana collection (Zoraka)<br />

etc (so that R371a*** is from the same collection event and<br />

spatial unit as R372a***, R373a*** etc.)<br />

Sherds prefixed P, C, A, B, GH, GE, GW, MH, HG99, K, and NR were from single-unit<br />

collections. For these sherds the remaining digits indicate sherd identity within the unit.<br />

Sherds prefixed Z, HV and HG have a collection unit number following the prefix, as<br />

follows.<br />

Zangana: Zxxxyyy where xxx is a three digit unit identifier and yyy is the sherd<br />

number.<br />

Honiavasa: HV0xyyy where 0x is unit 01-05 and yyy is sherd identity<br />

Hoghoi :HGxxsyy or HGxxayy where ‘xx’ is the unit, ‘yy’ or ‘yyy’ is the sherd<br />

number, ‘s’ means subsurface sample and ‘a’ means additional collection<br />

(a control for additional time spent in collection, to prevent intensity <strong>of</strong><br />

collection bias in the sherd density pattern across the site). <strong>The</strong> usual form<br />

is HGxxyyy. Where multiple fabric readings were taken as a quality control<br />

measure there are additional records in the descriptive database with ‘b’,<br />

‘c’ or ‘a’ added into the unit number or sherd number in place <strong>of</strong> a zero.<br />

197


Sherd Size:<br />

Sherd size was recorded as sherd area, sherd weight and in the case <strong>of</strong> lip and carination<br />

sherds, estimated vessel equivalent (EVE) (Egl<strong>of</strong>f 1973, Orton 1982, 1993, Orton & Tyers<br />

1991). Sherd area was counted using a gridded PVC transparency, which could partially<br />

conform to sherd external curvature. Measurement was made on the exterior <strong>of</strong> sherds, as<br />

a count <strong>of</strong> square centimetres, with an accuracy estimated at ± 10%. Sherd weight was<br />

recorded to the nearest gram using Swedish rounding in general, although initially decimal<br />

places were recorded also. Date and time were input automatically by the database and<br />

were used to correct some temporal inconsistencies in data recording. An additional<br />

measure in this regard was the re-analysis <strong>of</strong> the Paniavile assemblage which had been the<br />

first assemblage analyzed. A “notes” field was used primarily in initial database<br />

development, to record additional information usually coded subsequently in added fields.<br />

Fabric/Temper:<br />

Fabric descriptions were initially made using a point-counting system, which proved too<br />

time-consuming; and also cluster analysis on the results was inconsistent with sherd<br />

groupings established by visual comparisons. This initial approach was discarded. Fabric<br />

grouping <strong>of</strong> sherds was subsequently done using iterative manual sorting and visual<br />

comparison. A corner <strong>of</strong> each sherd was snapped, and most sherds had previously been<br />

acid-etched in the process <strong>of</strong> cleaning marine encrustation. Examination <strong>of</strong> etched sherd<br />

surfaces and the fresh un-etched break allowed reliable discrimination between calcite<br />

fragments, quartz/feldspar fragments, volcanic rock fragments, and Ferromagnesian<br />

mineral fragments, when examined under 10x binocular magnification.<br />

Iterative comparisons were used to develop a fabric type-series, but this also<br />

proved overly time-consuming. An ordinal notational system describing type and size <strong>of</strong><br />

inclusions in declining order <strong>of</strong> abundance was finally used. <strong>The</strong> approach is convergent<br />

with that <strong>of</strong> Wickler (Wickler 2001: 97-99). Major classes <strong>of</strong> grain identifiable under<br />

binocular microscope are listed in Table 8.<br />

198


Table 8: Classes <strong>of</strong> mineral identified at 10x magnification in reflected light.<br />

v ferromagnesian minerals ( opaque ferromagnesian, pyroxene, olivene,<br />

hornblende) black or green<br />

c calcite reef detritus, white/grey, s<strong>of</strong>t, <strong>of</strong>ten partially sintered and easily<br />

acid-etched<br />

q hard translucent mineral grains: quartz or feldspar<br />

t polycrystalline lithic fragments, usually volcanic, black-grey and become<br />

light grey-silver when etched<br />

w white volcanic glass fragments<br />

r red volcanic glass fragments<br />

b black volcanic glass fragments<br />

A microscope eyepiece was fitted with a 10x10 grid <strong>of</strong> 0.5mm squares to aid grain size<br />

estimation. Size descriptor suffixes for the mineral groups were as follows: x (poorly<br />

sorted sand incorporating a range <strong>of</strong> sand-sized grains from 0.1mm to 2mm), l (large), m<br />

(medium), s (small) and f (very fine). <strong>The</strong>se size classes corresponded to >0.5mm, 0.49-<br />

0.25mm, 0.24-0.1mm and


Table 9: Examples <strong>of</strong> descriptive syntax for tempers.<br />

VXCS Volcanic-Calcite hybrid temper comprising dominant Volcanic<br />

ferromagnesian minerals <strong>of</strong> miXed size range with subordinate Calcite<br />

(Small) reefal detritus grains<br />

CSQM Quartz-Calcite hybrid temper comprising Calcite reef detritus (Small)<br />

dominant, with subordinate Quartz/feldspar grains (Medium-sized)<br />

VFSMCF Volcanic particles (very Fine, probably opaques) dominant with a few<br />

Small sized and occasional Medium-sized volcanic particles,<br />

subordinate Calcite grains (very Fine)<br />

established for sites other than Miho and Honiavasa, and for the “P***” sample from<br />

Paniavile, which was etched. <strong>The</strong>se two Groups (1 & 3) while distinct in the sample sent<br />

to Dickinson, form something <strong>of</strong> a continuum, where some sherds are intermediate<br />

between the types, suggesting a more moderate degree <strong>of</strong> placer effect than seen in the<br />

placer volcanic samples. Groups 1 and 3 are therefore regarded as subgroups <strong>of</strong> a single<br />

class.<br />

Group 2, un-placered hornblendic-feldspathic volcanic sand tempers, include sherds<br />

<strong>of</strong> similar unplacered feldspathic composition, but the relative abundance <strong>of</strong> hornblende is<br />

thought to be highly variable within this group. <strong>The</strong> sherds examined by Dickinson and<br />

assigned to this group were all relatively rich in hornblende, and were interpreted as most<br />

likely from Vella Lavella. Many sherds in this group have pyroxene rather than hornblende<br />

as the predominant ferromagnesian mineral, as far as could be established from<br />

examination <strong>of</strong> sherd surfaces under 10x binocular magnification, so this group should be<br />

regarded as comprising a particularly low level <strong>of</strong> placering, consistent with stream sands,<br />

rather than as being entirely exotic. Some sherds within this grouping might be exotic to<br />

the New Georgia Island Group, but many could turn out to be local were further<br />

petrographic analysis carried out.<br />

200


Low-power reflected-light discrimination between Groups 4, 5, and 6 was problematic in<br />

Table 10: Temper groupings after Dickinson 2000.<br />

Group 1 Ferromagnesian Placer Volcanic sand tempers<br />

Group 2 Unplacered Hornblendic-Feldspathic Volcanic Sand Tempers<br />

Group 3 Pyroxenic-Lithic Volcanic Sand Tempers<br />

Group 4 Quartz-Bearing Volcanic Sand Tempers<br />

Group 5 Quartz-Calcite Hybrid Sand Tempers<br />

Group 6 Calcite Tempers (some with possible subordinate fine ferromagnesian<br />

component)<br />

Group 7<br />

(Group 1?)<br />

Placered volcanic with plagioclase-rich microlitic rock fragments<br />

Group 8 Temper not recorded<br />

many cases. While there were many examples <strong>of</strong> a distinctively quartz-calcite hybrid<br />

temper (Group 5), there were other sherds with a greater proportion <strong>of</strong> volcanic fragments,<br />

which tended towards the volcanic-quartz grouping (Group 4). Also, there were quartz-<br />

calcite sherds with very low amounts <strong>of</strong> quartz, difficult to detect in sherd surfaces under<br />

binocular magnification, but which, when analyzed petrographically, turned out to have a<br />

terrigenous fraction <strong>of</strong> similar composition to the quartz-calcite sherds (see Appendix 1:<br />

Table 196-5) It seems clear that varying proportions <strong>of</strong> terrigenous component to calcite<br />

reef detritus component can account for such a variety <strong>of</strong> visual temper characteristics yet<br />

allow derivation <strong>of</strong> these materials from a single coastline (Felgate & Dickinson 2001).<br />

Equally however, it is possible that some <strong>of</strong> the calcite-tempered sherds for which a<br />

quartzo-feldspathic fraction cannot be discerned in the sherd surface could be from a non-<br />

quartzose source area. Further petrographic quality-control is necessary to assess the<br />

situation in this regard between Groups 5 and 6.<br />

Group 7 in the database refers to a fabric in which placered volcanic temper is<br />

dominant, but in which subordinate dark grey microlitic rock fragments occur. Due to<br />

201


constraints on sample size, none <strong>of</strong> these sherds were included in the sample analyzed by<br />

Dickinson. <strong>The</strong>se are provisionally considered to be a less placered variant <strong>of</strong> the<br />

ferromagnesian placered volcanic group (Group 1) with affinities also to pyroxenic-lithic<br />

volcanic sand tempers (Group 3) and thus intermediate between Groups 1 and 3, although<br />

calcite grains were rare or absent in Group 7, suggesting a separate origin.<br />

Group 8 comprises those sherds for which fabric was not recorded, usually due to<br />

difficulties discerning temper type under binocular magnification, most commonly as a<br />

result <strong>of</strong> a dark reduction core which hid darker-coloured mineral grains in reflected light.<br />

Vessel Parts:<br />

Definition <strong>of</strong> vessel part terminology is fundamental to the ceramic data structure (Figure<br />

29, Figure 30). In some instances these vessel part definitions seem more specific than<br />

those <strong>of</strong> potters in the past, as will be argued in Chapter 9. Identification <strong>of</strong> vessel part for<br />

smaller sherds was not always possible, and sometimes a default assignment was used<br />

when a sherd met some criterion. For example all sherds <strong>of</strong> hyperboloid form (Rice<br />

1987:219) other than those from Honiavasa conical-shouldered carinated jars were<br />

classified as rims.<br />

Thickness, Form and Decoration <strong>of</strong> Various Vessel Parts Represented on Sherds:<br />

Three tables containing information on thickness, form and decoration were held in a one-<br />

to-many relationship with the master table via the sherd ID field. This allowed information<br />

pertaining to multiple specific parts <strong>of</strong> the vessel to be recorded separately, yet as a group<br />

for that sherd. By way <strong>of</strong> example, for a sherd where the lip, rim and neck were<br />

represented, the thickness table could contain measurements taken according to a set <strong>of</strong><br />

conventions at each <strong>of</strong> these points on the vessel, allowing thickness trends to be<br />

calculated for that sherd, and also allowing comparison <strong>of</strong> assemblages in terms <strong>of</strong><br />

location-controlled thickness measurements. Similarly, for vessel form, characteristics <strong>of</strong><br />

the lip, rim, neck, shoulder, carination and body could be recorded where present for each<br />

sherd, allowing virtual reconstruction <strong>of</strong> sherd form from the database, and also allowing<br />

202


Figure 30: Major vessel form variants showing part<br />

terminology; L=lip, R=rim, N=neck, S=shoulder,<br />

C=carination, B=body; inverted or unrestricted vessels have<br />

no neck, while for restricted vessels with everted rims (the<br />

first seven) the only distinction in neck types is between the<br />

double neck (top centre) and the single neck (including all<br />

unlabelled).<br />

detailed comparison <strong>of</strong> assemblage variability controlling for position on the vessel.<br />

Decoration was recorded in the same way, i.e. the decorative characteristics <strong>of</strong> each vessel<br />

part represented on the sherd were recorded in some detail. <strong>The</strong>re could be more than one<br />

decoration record for one part <strong>of</strong> a sherd, provided that the decorative technique differed,<br />

as the key for the decoration table was a composite <strong>of</strong> more than one column.<br />

A similar effect could have been achieved using a single flat table <strong>of</strong> data with a<br />

203


large number <strong>of</strong> attributes or variables, but this would have been unwieldly for data entry,<br />

requiring a multi page electronic data entry form to achieve the same level <strong>of</strong> detail as<br />

entered in the relational system.<br />

Thickness: For the table <strong>of</strong> sherd thicknesses by vessel part, the “ID” field linked<br />

to the master table record for that sherd (Table 11). “Part” could be either the lip (l),<br />

rim(r), neck(n), shoulder(s), carination(c), body(b), base(a) or unknown(u). “Part”<br />

comprised a secondary index, to enable only one measurement for each part, and to give<br />

each part measurement a unique identity. Thickness was measured according to the rule<br />

for that part, using digital calipers and measured in mm, using Swedish rounding in most<br />

cases, although some initial measurements recorded thickness to one decimal place.<br />

Table 11: Data structure <strong>of</strong> table <strong>of</strong> thicknesses for<br />

each part <strong>of</strong> the sherd.<br />

Field Name Type Size Key Required Value<br />

Id A 7 Yes<br />

Part A 1 Yes Yes<br />

Thickness N Yes<br />

Thickness at the lip was measured perpendicular to the tangent <strong>of</strong> the interior surface <strong>of</strong><br />

the rim where it meets the lip. Thickness <strong>of</strong> the rim was measured equidistant between the<br />

lip and the point <strong>of</strong> maximum vertical curvature <strong>of</strong> the neck (minimum vertical radius). In<br />

the case <strong>of</strong> broken rims, where the full depth was not present, thickness was measured at<br />

the lowest point. Thickness measurements were an average <strong>of</strong> the reading from both ends<br />

<strong>of</strong> the sherd (in the horizontal dimension) where the sherd could be oriented. For sherds<br />

<strong>of</strong> unknown part, which could not be oriented, thickness was recorded as the average <strong>of</strong><br />

several caliper readings across the sherd.<br />

204


On the data entry form for the sherd, on which the thickness table resided, once a part had<br />

been entered, a thickness for that part was required by the database, and use <strong>of</strong> the<br />

secondary index meant that only one record per part could be entered.<br />

Vessel Form:<br />

Vessel forms were recorded using interval variables as far as possible (the set <strong>of</strong> curve-<br />

fitting templates used provided the intervals). Secondary classification was undertaken<br />

when these observation suggested essential classes in the data as opposed to continuous<br />

variation. While some vessel form descriptive schemes measure the orientation <strong>of</strong> the rim<br />

and the angle between neck and shoulder in some manner (see Irwin 1972:60: inclination<br />

angle), the effects <strong>of</strong> both assemblage brokenness and rim/neck/shoulder curvature on the<br />

measured angle are seldom discussed. Also, necks tend to be treated as though they are<br />

the apex <strong>of</strong> two straight lines, these lines being the vertical sections <strong>of</strong> the rim and<br />

shoulder. In reality, these sections are made up usually <strong>of</strong> a series <strong>of</strong> complex curves rather<br />

than straight lines and corners.<br />

<strong>The</strong> approach taken in this study was to regard these complex curves as reduceable<br />

not to a series <strong>of</strong> straight lines <strong>of</strong> measured length and angles, but to a series <strong>of</strong> simple arcs<br />

<strong>of</strong> circles <strong>of</strong> measurable length, oriented by measurable angles in relation to each other.<br />

In my opinion this approach provides a better approximation <strong>of</strong> the actual shape <strong>of</strong><br />

ceramic vessels in vertical pr<strong>of</strong>ile, and therefore does a better job <strong>of</strong> describing variation<br />

in vessel form. This approach is less sophisticated than the two-curvature method <strong>of</strong><br />

Hagstrum and Hildebrand, where the exterior vertical section can be represented as a<br />

continuously varying curve (Hagstrum & Hildebrand 1990), but at the level <strong>of</strong> sherd<br />

assemblage description there is substantial convergence between my approach and theirs.<br />

Curvature measurements were taken at various parts <strong>of</strong> the vessel as represented by<br />

sherds, and covariation between interval attributes was assessed by means <strong>of</strong> bivariate<br />

plots.<br />

205


Field<br />

Name<br />

Type Size Key _Require<br />

d Value<br />

Id A 7 Yes Yes<br />

Part A 1 Yes Yes<br />

Type A 1<br />

Angle N<br />

Vcurve N<br />

Hcurve N<br />

Sherdfor<br />

m notes<br />

Curvatures in all cases were measured using brass radius templates, manufactured<br />

by the <strong>Auckland</strong> <strong>University</strong> Engineering Workshop for the Anthropology Department.<br />

Internal radii <strong>of</strong> the templates, used to measure the horizontal curvature <strong>of</strong> the exterior <strong>of</strong><br />

the vessel (including lip Hcurve), ranged from 50 to 300 mm in increments <strong>of</strong> 10mm (5mm<br />

increments were manufactured but provided spurious accuracy given the hand-formed,<br />

irregular nature <strong>of</strong> the pottery). <strong>The</strong> brass templates had a width <strong>of</strong> 7mm, so external radii<br />

measurable by the set <strong>of</strong> (concave) curves ranged from 57 to 307 mm, also in 10mm<br />

increments as used.<br />

Data structure for vessel form (Table 12) used “Id” and “part” information as for<br />

the thickness table. How the remaining fields were filled out depended on both the part and<br />

the part form type (the first column is the list <strong>of</strong> data fields which form the columns in the<br />

data table, the other columns detail properties <strong>of</strong> each data field).<br />

Variations in the way different vessel part forms were recorded are listed below<br />

Table 12: Data structure for the table <strong>of</strong> records <strong>of</strong> sherd form attributes by vessel part.<br />

A 240<br />

206<br />

_Min<br />

Value<br />

_Max<br />

Value<br />

Rimdepth N 0.00 200.00


under sub-headings:<br />

Lip Form:<br />

For lips, nine lip “type” options were used (Figure 31), with remaining fields filled out the<br />

same regardless <strong>of</strong> type. Lip angle was measured to the nearest ten degrees using a<br />

goniometer, as the angle between the uppermost tangent to the interior surface <strong>of</strong> the rim,<br />

and the innermost facet <strong>of</strong> the lip, except in the case <strong>of</strong> rounded lips, for which lip angle<br />

was not measurable. Measuring this angle, and another which gave the degree <strong>of</strong> rim<br />

eversion, allowed a reduction in the number <strong>of</strong> lip types recorded below the norm for<br />

Pacific potsherd studies, without loss <strong>of</strong> information on variability.<br />

<strong>The</strong> “Vcurve” field (Vertical curvature) was not used for describing lip form. <strong>The</strong><br />

“Hcurve” field (Horizontal curvature) was measured by curve fitting for 45 lips, these<br />

tending to be the larger sherds where the measurement is less likely to be spurious. Of the<br />

Figure 31: Lip form variants showing<br />

database codes<br />

207


42 sherds in this category for which EVE was recorded, EVE ranged from 3% to 17%,<br />

with an average value <strong>of</strong> 7.2%. A larger number <strong>of</strong> lip sherds were measured in this way<br />

during preparation <strong>of</strong> sherd drawings, where multiple measurement points were used to<br />

establish the best location <strong>of</strong> the central vertical axis (CVA) <strong>of</strong> vessels, but these<br />

measurements, while shown on some sherd illustration figures, were not entered into the<br />

database. In the drawings, irregular Hcurves which did not accurately conform to any <strong>of</strong><br />

the curve templates were shown as a dashed CVA estimate, while solid lines indicated a<br />

good circular fit.<br />

“Sherdform notes” was a general purpose field used particularly in the development<br />

stages <strong>of</strong> database design to note any data that required systematization. “Rim Depth” was<br />

a general purpose field applicable to rims, some types <strong>of</strong> neck, and shoulders, but was not<br />

used for lip data.<br />

Rim Form:<br />

Sherd ID and “part” were used as for lips. Eight rim types were distinguished (Figure 32),<br />

and measurement rules for rims varied slightly with rim type in some cases. Generally<br />

speaking, the rim descriptive terminology <strong>of</strong> Snow was used (Snow & Shutler 1985:20)<br />

which is similar to that used by Irwin (Irwin 1972:55, 1985). Rims which diverge with<br />

height from the CVA <strong>of</strong> the vessel are referred to as everted. Rims which approach the<br />

CVA with increasing height are inverted, while those which do neither are vertical. Snow<br />

and Shutler provide a finer descriptive terminology that subdivides everted rims into<br />

straight everted, excurvate everted, or flared everted, excurvate having a concave exterior<br />

208


vertical contour (hyperboloid form) and flared being convex in that plane. <strong>The</strong> “type” field<br />

was sufficient to encompass all these variants, in addition to some sub-types <strong>of</strong> everted<br />

rims. Eight rim types were defined (Figure 32). Rim pr<strong>of</strong>ile (Snow & Shutler 1985:22)<br />

was not recorded, as thickness measurements for lip, rim and neck captured this variation.<br />

For excurvate rims (x) or straight rims (s), where the lip was present, rim angle<br />

was recorded as the angle between the CVA and the interior tangent (in vertical contour)<br />

at the lip. Where local pr<strong>of</strong>ile concavity was present through thickening <strong>of</strong> the lip, the<br />

goniometer would sit stable on two points, while on convex-pr<strong>of</strong>ile interiors, the<br />

Figure 32: Rim types (database codes shown and<br />

measurement method for rim angles).<br />

209


goniometer would be positioned so that it contacted as a tangent below the lip, and was<br />

then rocked outwards until the point <strong>of</strong> contact reached the top <strong>of</strong> the rim (see “rim slope”<br />

in Smith 1985:263). For flared everted rims (f), the goniometer could be rested stable<br />

across two contact points, one at the neck and one at the rim.<br />

This process was not particularly accurate, but had the advantage <strong>of</strong> being<br />

measurable independent <strong>of</strong> the neck (except in the case <strong>of</strong> flared rims), and required only<br />

that there was enough <strong>of</strong> the lip present to allow the sherd to be oriented, itself a<br />

procedure fraught with error, particularly with uneven hand-formed pottery. This meant<br />

a larger sample <strong>of</strong> measurements could be obtained, as sherds having both the lip and the<br />

neck present were rare. This measurement is similar to the more commonly measured<br />

“orientation angle” (see for example Irwin 1972:60) but not the same, except in the case<br />

<strong>of</strong> flared or straight rims. Where the rim is excurvate this rim angle measurement will be<br />

substantially wider than Irwin’s measure <strong>of</strong> orientation angle, dependent on the amount <strong>of</strong><br />

excurvature as expressed by the Vcurve measurement (see below) and the depth or height<br />

<strong>of</strong> the rim.<br />

For excurvate everted rims (x), brass curvature templates were used to determine<br />

the best expression <strong>of</strong> concavity in the vertical plane (Rim Vcurve, Figure 33), recorded<br />

as radii in millimetres, in increments <strong>of</strong> 10mm. <strong>The</strong> smallest curve template was 47mm,<br />

curves <strong>of</strong> smaller radius were estimated using a ruler or gridded transparency. For Flared<br />

rims (f) the curve measurement indicates the radius <strong>of</strong> concavity in the vertical plane. If<br />

both lip and neck were present, horizontal curvature (Hcurve) was sometimes recorded<br />

210


midway between lip and neck, on the exterior surface. No analytical use was made <strong>of</strong> this<br />

measurement in the present study. Rim depth (Figure 33) was measured in millimetres<br />

with calipers between the outer edge <strong>of</strong> the lip and the topmost point <strong>of</strong> maximum<br />

curvature (minimum radius) <strong>of</strong> the neck, where the Vcurve template would be tangential<br />

to the Vcurve template <strong>of</strong> the neck (see below). Where either the neck or the lip was not<br />

present rim depth was noted as greater than the measured distance to the broken vertical<br />

boundary <strong>of</strong> the rim, and recorded as a “Sherdform note” rather than in the “Rimdepth”<br />

field.<br />

Taken together with the lip and rim thickness measurement, these approximately<br />

reconstruct the form <strong>of</strong> rims, and therefore allowed a more realistic record <strong>of</strong> rim form<br />

variability, more likely to capture subtle variation than previous schemes, while boosting<br />

sample size. As discussed in the preamble to the ceramic analysis, this level <strong>of</strong> detail<br />

seemed desirable in attempting to characterised stylistically depauperate assemblages, and<br />

to analyse variability in detail. For more complex rim forms, hard flare (h), compound(c),<br />

hard compound (k) or modeled (m), measurements were made in the same way, but on the<br />

upper portion <strong>of</strong> the rim only, so that measurements could be obtained for more broken<br />

examples where the neck was not present. In such cases rim depth too was measured from<br />

the outer edge <strong>of</strong> the lip to the first corner point rather than to the neck.<br />

In practice, the advantage <strong>of</strong> being able to measure rim angle independent <strong>of</strong> the<br />

neck was mitigated slightly by the difficulty <strong>of</strong> orienting more broken sherds which did not<br />

include the neck. For some <strong>of</strong> the slab-built Honiavasa pottery though, which tended<br />

211


to have broken at the neck, this system allowed more measurements than it would have<br />

if Irwin’s or Snow’s system had been used.<br />

Neck Form:<br />

<strong>The</strong> vessel neck in this study was defined as the orifice restriction between body and rim<br />

(Figure 30). Unrestricted vessels <strong>of</strong> incurvate or direct form (Snow & Shutler 1985:20)<br />

therefore do not have a neck. <strong>The</strong> double neck vessel form (d) (Figure 30) is the only neck<br />

type with depth, and for which neck depth must be recorded to record the shape <strong>of</strong> the<br />

vessel. In double necks an upper and lower tight radius are separated by a distance, over<br />

which the vertical external section is roughly straight. Other neck types are restricted (r)<br />

or unrestricted (u), meaning there is no neck or a slight inflection (indirectness) in the<br />

external pr<strong>of</strong>ile. <strong>The</strong>se are described only in terms <strong>of</strong> neck type, angle (Figure 33), Vcurve<br />

(Figure 33) and Hcurve. <strong>The</strong> neck seems to have been a rugged portion <strong>of</strong> the<br />

Figure 33: Measurement <strong>of</strong> rim depth, rim Vcurve, neck<br />

angle and neck Vcurve at different levels <strong>of</strong> brokenness.<br />

212


vessel, well represented in the sherd assemblages in most <strong>of</strong> the site samples, and<br />

conveying considerable information on vessel form when thus described.<br />

Neck Hcurve was measured on the external surface <strong>of</strong> the vessel. Where sherds<br />

were smaller than 10cm 2 , or Hcurve was too irregular to fit any <strong>of</strong> the templates well,<br />

measurement was either not taken or disregarded by use <strong>of</strong> a sherd size filter during<br />

analysis. Orifice radii can be obtained by subtracting neck thickness from neck Hcurve.<br />

Neck “Vcurve” was estimated using templates for curves <strong>of</strong> radius $47mm, and<br />

by estimation for smaller curves. <strong>The</strong> neck Vcurve was the largest circle that could fit<br />

vertically and snugly against the neck in the middle <strong>of</strong> the sherd (Figure 33). Minor<br />

irregularities might cause light to show in places between the neck and the template, but<br />

usually this measurement was a straightforward and unambiguous procedure.<br />

<strong>The</strong> neck Vcurve measurement forms the basis for the neck angle measurement<br />

Figure 34: Shoulder form measurement.<br />

213


shown in Figure 33. <strong>The</strong> advantage <strong>of</strong> this approach is that the neck angle is measurable<br />

at the neck, rather than requiring the lip to be present, as is more common (e.g. Irwin<br />

1972:60). This means that for a relatively broken assemblage where few lip-rim-neck<br />

sherds are present, neck angle measurements can still be obtained in quantity, yielding a<br />

valuable increase in the sample size for this type <strong>of</strong> information on vessel form for the<br />

assemblage. Also, this method does not treat the neck as though it is an “inflection point<br />

(emphasis added)” (Wickler 2001:93), but rather as a curve, which is a less arbitrary<br />

determination <strong>of</strong> where the measurement should be taken, particularly where the neck<br />

Vcurve is gentle.<br />

Shoulder Form:<br />

Four shoulder types were defined (Figure 34). For s<strong>of</strong>t shoulders (s), the most numerous<br />

category, shoulder, Vcurve and Hcurve radii were defined as for necks, with one important<br />

difference: the templates used to measure curves were too long to be <strong>of</strong> much use<br />

measuring tight convex curves on shoulders, and Vcurve and shoulder angle were thus not<br />

measured as accurately as in the case <strong>of</strong> necks. S<strong>of</strong>t shoulder angle measurements on small<br />

sherds (


characteristics. Shoulder angle for hard shoulders was effectively the angle between the<br />

tangents to the upper and lower halves <strong>of</strong> the shoulder at the apex <strong>of</strong> the halves. Notes<br />

referring to bevel distance on these sherds are the depth measurement, recorded also in the<br />

Rimdepth field.<br />

For slab-built vessels where a lip-rim-neck slab is joined directly to a shallow body<br />

(Figure 32:“k”), there was no convex (in external pr<strong>of</strong>ile) formed shoulder and these were<br />

therefore classified as shoulderless (l) type, where vcurve measures the concavity <strong>of</strong> the<br />

external pr<strong>of</strong>ile from the lower boundary <strong>of</strong> the neck to the carination/modeling junction<br />

edge. Hcurve was either not measured or disregarded in subsequent analysis.<br />

Where the shoulder was straight in external pr<strong>of</strong>ile or gently convex, and typically<br />

extended, the shoulder type was designated conical ( c ), and was expected to have, or did,<br />

join to the maximum width <strong>of</strong> the vessel. In these cases Vcurve measured the degree <strong>of</strong><br />

convexity <strong>of</strong> this flattish shoulder, and the distance from the lower boundary to the neck<br />

to the beginning <strong>of</strong> the carination was recorded in the “Rimdepth” field.<br />

Carination Form:<br />

Carinations could be either hard (h) (small Vcurve) or s<strong>of</strong>t (s) (larger Vcurve) or double<br />

(d) (there were two examples: see Figure 9 for one <strong>of</strong> these). Carination angle was<br />

measured by the same rules as neck angle or shoulder angle. Carination Hcurve was<br />

measured on the exterior plan <strong>of</strong> the vessel (horizon).<br />

Body Form:<br />

Body sherds were all convex in form by definition, and were classified as either <strong>of</strong> conical<br />

215


type (c), where orientation could be established and where the vertical curvature was slight<br />

in comparison to the horizontal curvature, or <strong>of</strong> globular (spheroidal) type (g) where both<br />

dimensions were similar (Figure 35). Latitudinal rows <strong>of</strong> anvil marks, and horizontal<br />

wiping marks were useful in inferring body sherd orientation in many cases. In cases where<br />

the sherd could not be oriented, maximum and minimum curvature were entered in the<br />

notes field <strong>of</strong> the form table, rather than in Hcurve and Vcurve fields, unless maxima and<br />

minima were the same (spherical form). Where body sherds were smaller than 10cm 2 ,<br />

Vcurve and Hcurve data were regarded as deriving from an unknown vessel part,<br />

irrespective <strong>of</strong> coded vessel part, as these could equally be portions <strong>of</strong> rounded shoulders<br />

rather than body sherds.<br />

For those sherds that could not be oriented, but could be characterized as conical<br />

(C) or globular/spheroidal (G), a number <strong>of</strong> body forms are possible, and some confusion<br />

with conical shoulder forms is possible also. Relative abundance <strong>of</strong> various body sherd<br />

Figure 35: Derivation <strong>of</strong> conical/cannister ( C ) and spheroidal (G) sherds from<br />

various body forms.<br />

216


forms in assemblages must be interpreted with these various possibilities in mind, and may<br />

serve to provide a basis for establishing difference between assemblages, but not similarity.<br />

Interior Form:<br />

It was noticed during analysis that some interiors were smooth, with even body wall<br />

thickness, while others had anvil marks arranged <strong>of</strong>ten in latitudinal circles around the pot.<br />

This was thought to relate to whether or not a final or secondary smoothing had been<br />

performed, using a larger anvil than that used to rough out the body shape. It seemed also<br />

that there was some correlation between this secondary smoothing and a more distinct, or<br />

sharper edge to the interior <strong>of</strong> the neck (orifice). To test whether this<br />

correlation was real, the larger sherds were re-analyzed to record whether the body<br />

interior was even (e) (Figure 36) or uneven, the latter involving the following options:<br />

finger impressions (f), anvil impressions (a) or other (o) specified in the notes field. Also<br />

recorded was whether the interior <strong>of</strong> the neck was sharp or rounded. <strong>The</strong>se data were<br />

subsequently filtered to exclude small sherds from analysis, to preclude the possibility <strong>of</strong><br />

Figure 36: Form codes for vessel interiors.<br />

217


misinterpretation based on too small a sample <strong>of</strong> the vessel.<br />

Unknown Part Form:<br />

<strong>The</strong>se were probably mostly s<strong>of</strong>t shoulders or body sherds, most being nondescript<br />

fragments <strong>of</strong> more-or-less globular form. Where there was some basis for orienting the<br />

sherd, Vcurve and Hcurve were recorded. Descriptive statistics <strong>of</strong> these measurements,<br />

filtered to sherds over 10cm 2 in area were subsequently used in analysis. Some sherds<br />

recorded as rim sherds on the basis <strong>of</strong> their hyperboloid form (concave in one surface axis<br />

and convex in the other) could be parts <strong>of</strong> shoulderless vessels below the neck, but this<br />

seems unlikely, as all decorative evidence points to these being rims. Shoulders <strong>of</strong> that<br />

form were all decorated with bounded incision, except for one vessel from the Paniavile<br />

site which had idiosyncratic decoration, and one plain carination.<br />

Base Form:<br />

Very few base sherds were identified, and those tentatively. No evidence was found for flat<br />

bases, and only tenuous suggestion <strong>of</strong> some more conical forms, and this is regarded<br />

Figure 37: Possible conical base sherds from robust vessels.<br />

218


as indirect evidence for predominantly rounded bases in the entire assemblage. Base sherds<br />

were identified through having a parabolic rather than circular curvature in any dimension<br />

around a point (see the base pr<strong>of</strong>ile <strong>of</strong> the bi-conical form in Figure 35). One or two<br />

sherds have slight flattening and extreme thickness also (see Figure 37 for illustrated<br />

examples).<br />

Decoration:<br />

Objectives <strong>of</strong> the decorative study shifted significantly through the course <strong>of</strong> data recording<br />

and analysis, from an initial poorly-defined aim <strong>of</strong> a detailed social/behavioural analysis,<br />

to a position where understanding production variability was seen as the key to sample<br />

evaluation (allowing the construction <strong>of</strong> vessel families) and to constructing a seriation<br />

chronology. This shift was a consequence <strong>of</strong> the realization that the sherd sample displayed<br />

a low degree <strong>of</strong> vessel completeness for all collection sites, and must have been heavily<br />

sampled from a breakage population by post-discard processes (Felgate 2002). Nowhere<br />

in the analysis is this shift more apparent than in the analytical scheme used to describe<br />

surface decoration <strong>of</strong> the pottery. Initially, I was interested in identifying the work <strong>of</strong><br />

individual potters through examination <strong>of</strong> variation in technical execution, but it became<br />

apparent, as will be discussed, that there was no particular reason to expect a good sample<br />

<strong>of</strong> the work <strong>of</strong> a single potter to have been preserved in the assemblages, and the primary<br />

focus <strong>of</strong> research shifted from the fine details <strong>of</strong> decoration to a slightly broader analysis<br />

<strong>of</strong> production variability. <strong>The</strong>re is substantial data-redundancy in the decorative<br />

information recorded as a result <strong>of</strong> this shift, although some use was made <strong>of</strong> the mark<br />

variability data in testing units <strong>of</strong> decorative classification (particularly for<br />

219


impressed or notched lips).<br />

Data structure for the decoration table are shown in Table 13. Use <strong>of</strong> the fields<br />

“Id”, “Part” and “Type” were as for the form table <strong>of</strong> description, with multiple records<br />

for sherds linked to the master table <strong>of</strong> sherd properties via the “Id” field. <strong>The</strong>se fields, and<br />

Table 13: Data structure for decoration records.<br />

Field Name Type Size Key Required<br />

Value<br />

Id A 7 * *<br />

Part A 1 * *<br />

Type A 1 *<br />

Dec_tech A 3 * *<br />

Pattern A 3<br />

Pattern_rep_dist N<br />

Notes_1 A 240<br />

Element A 3<br />

ERD N<br />

Ms_MarkSection A 1<br />

Ml_MarkLength N<br />

Mw_MarkWidth N<br />

Md_Markdepth N<br />

Hand A 5<br />

Notes A 240<br />

LinDecClass S<br />

LipDecCode S<br />

NeckDecCode S<br />

a fourth, decorative technique (“Dec_tech”) were hierarchically indexed in the database<br />

to allow each vessel part to have more than one decorative technique recorded. In practice<br />

it was found more convenient to have multiple decorative techniques on the same part<br />

recorded as coded combinations rather as separate lines in the table, in which<br />

220


Table 14: Decorative techniques.<br />

Code Description<br />

? Decorative technique uncertain<br />

a Applique (Figure 38, Figure 39, Figure 40)<br />

b Burnishing or polishing (not illustrated: not used in analysis)<br />

d Lip deformation (Figure 41, Figure 42, Figure 43)<br />

e Excision (Figure 44, Figure 45)<br />

f Fingernail impression (subdivided by the “element” field as<br />

detailed below)<br />

g Incision<br />

h Brushing (rare, identified on the presence <strong>of</strong> multiple parallel<br />

bristle marks sharing a common termination)<br />

i Impression not identified as fingernail impression, usually on<br />

lip or carination.<br />

l Linear stamping (rare, but see Figure 12: Z.10.573, which<br />

has linear stamping on the shoulder, in a pattern <strong>of</strong> opposed<br />

lines forming triangles)<br />

m Spatulate tool impression (during vessel forming)?(see Figure<br />

46)<br />

o Other unclassified techniques (see “Notes_1")<br />

p Punctation (Figure 15 )<br />

r Excision by punctation/rotation (see Figure 47)<br />

s Surface smoothing without striation (self slip?)<br />

t Perforation (see Figure 48)<br />

v Anvil impression, exterior (unimportant- one unconfirmed<br />

example)<br />

w Wiping - identified by presence <strong>of</strong> multiple shallow rounded<br />

parallel grooves.<br />

y Wavy stamping or dentate stamping as specified in element<br />

field (Figure 45 and Figure 49)<br />

cases the predominant technique was described in detail in subsequent fields, and sub-<br />

techniques for that vessel part were described in the notes fields <strong>of</strong> the decoration table.<br />

<strong>The</strong> fields <strong>of</strong> decorative information were hierarchically structured from the most general<br />

to the most detailed. Decorative techniques were coded as in Table 14.<br />

221


“Pattern” (Table 13) is roughly equivalent to Anson’s concept <strong>of</strong> motif for more<br />

complex patterns, (Anson 1983:57) or Mead’s concept <strong>of</strong> “all<strong>of</strong>orm” (Mead 1975:23),<br />

being defined as the structure or arrangement <strong>of</strong> decorative elements, and is defined<br />

independent <strong>of</strong> decorative technique. Mead did not separate technique <strong>of</strong> execution and<br />

the pattern in which any technique was applied, thus generating unnecessary proliferation<br />

<strong>of</strong> some units <strong>of</strong> decoration, such as the various “Zone Markers” (see for example Mead<br />

1975:Figure 2.11 page 26, GZ2-GZ4). In this study such a distinction has been maintained<br />

as far as possible, with the objective <strong>of</strong> reducing the number <strong>of</strong> decorative classes.<br />

Pattern entries in the decoration table fell into two classes, formulaic patterns, from<br />

which the pattern could be generated by a formulaic repetition <strong>of</strong> elements, and more<br />

complex patterns described using sherd illustrations. <strong>The</strong> formulaic patterns are mostly<br />

simple latitudinal bands <strong>of</strong> decoration (pattern prefix “b” in the field “patt” in the<br />

decoration data table) running around the vessel parallel to the vessel horizon (with the<br />

central axis <strong>of</strong> symmetry oriented vertically). Pattern definitions are given in Table 15. <strong>The</strong><br />

complex patterns were later classified into two classes based primarily on zone markers,<br />

with a third dustbin class, mostly for pattern fragments.<br />

“Patt_rep_dist” (Pattern repeat distance - Table 13) was a measure (in mm) for simple<br />

formulaic patterns <strong>of</strong> the vertical distance between element repeats (from vertical centre<br />

to vertical centre). Thus in a pattern comprising multiple latitudinal bands <strong>of</strong> opposed pinch<br />

fingernail impression, the pattern repeat distance was a component in specifying the overall<br />

expression <strong>of</strong> the pattern on the sherd.<br />

222


Figure 38: Examples <strong>of</strong> applied decoration.<br />

Figure 39: Examples <strong>of</strong> applied decoration.<br />

223


Figure 40: Applied decoration on compound rims.<br />

Figure 41: Examples <strong>of</strong> deformation <strong>of</strong> the lip into a discontinuous band.<br />

224


Figure 42: Horizontal deformation <strong>of</strong> the lip into a continuous band in a wave pattern.<br />

Figure 43: Examples <strong>of</strong> discontinuous deformation.<br />

225


Figure 44: Excision <strong>of</strong> outer lip to form a band <strong>of</strong> notches.<br />

Figure 45: Compound rims from Nusa Roviana. NR.34 has excised lines forming the<br />

triangle pattern on the upper rim.<br />

226


Figure 46: Impressions on the top <strong>of</strong> the lip; and spatula impressions on the neck,<br />

thought to indicate forming <strong>of</strong> the neck using a tool.<br />

Figure 47: Excision by rotation: one end <strong>of</strong> a small twig or rod has been poked into the<br />

clay and the other end moved in a circle, leaving a conical hole.<br />

Figure 48: Perforation (upper hole).<br />

227


Figure 49: Examples <strong>of</strong> wavy stamping and an example <strong>of</strong> dentate-stamping.<br />

Figure 50: Applied decoration (in combination with incised decoration) with<br />

detachment scars indicating a v-pattern, and also possible attached disc.<br />

228


Table 15: Decoration pattern definitions<br />

“Patt” Description Anson Motif # Count<br />

? incomplete, insufficient to characterize 284<br />

?pi (continuous?) band <strong>of</strong> parallel elements on inner edge <strong>of</strong> part 1<br />

?po (continuous?) band <strong>of</strong> parallel elements on outer edge <strong>of</strong> part 1<br />

?pt (continuous?) band <strong>of</strong> parallel elements on top face <strong>of</strong> part 1<br />

acc accidental marks? 1<br />

am1 (complex applied pattern) see R37.11A.1 in Figure 40 2<br />

as1 (complex applied fillets) see R376a14 in Figure 38 1<br />

as3 see HV.1.3 in Figure 49 3<br />

as? (applied fillet ?) Not illustrated 1<br />

avc (applied v-shaped fillet) see Figure 50 1<br />

b continuous latitudinal band <strong>of</strong> an element (used for circular elements such as punctations or nubbins) 117<br />

bar incomplete, twinned vertical bars made up <strong>of</strong> fingernail impressions 1<br />

bd continuous latitudinal band <strong>of</strong> elements oriented diagonally to horizon (see Figure 51) 16<br />

bdi band, elements diagonally oriented, inner edge <strong>of</strong> lip (not illustrated) 4<br />

bdo band, elements diagonally oriented, outer edge <strong>of</strong> lip (equivalent to “bpi” below) 8<br />

bdt band, elements diagonally oriented, top face <strong>of</strong> lip (equivalent to “bpi” below, see Figure 57) 3<br />

bnd fragment <strong>of</strong> complex pattern? Repeats <strong>of</strong> bent lines, see MH266 (Figure 52) 272?, 273?, 498 sans zone markers? 3<br />

bo band, outer edge <strong>of</strong> lip (as for “bpi” below, but for circular elements) 6<br />

bot band <strong>of</strong> parallel elements in diagonally, symmetrically opposed sets, top face <strong>of</strong> lip (Figure 57) 1<br />

bp band <strong>of</strong> parallel elements, most commonly fingernail pinching but sometimes applied fillets (Figure 7) 124<br />

bpb bands, parallel elements, outer and inner edges <strong>of</strong> lip (Figure 19: Z.89.549, Figure 22, Figure 26: GE266) 16<br />

bpi band <strong>of</strong> parallel elements, inner edge <strong>of</strong> lip (Figure 54) 39<br />

bpo band <strong>of</strong> parallel elements, outer edge <strong>of</strong> lip (Figure 55, Figure 56) 69<br />

bpt band, parallel elements, top face <strong>of</strong> lip (Figure 57) 38<br />

bt band <strong>of</strong> elements on the top face <strong>of</strong> the part (applies to circular elements such as punctation) 4<br />

c10 complex linear pattern: see HV.4.164 (Figure 8) similar 154, 155 4<br />

c11 complex linear pattern: see HV.4.202, HV.2.302, HV.2.297 (Figure 8, Figure 9) 155 3<br />

c12 complex linear pattern: see HV.2.227 (Figure 49) 1<br />

c13 complex linear pattern: see HV.1.314 (Figure 9) 2<br />

c14 complex linear pattern: see HV.02.419 (Figure 58) 155 in fingernail impression? 4<br />

c17 complex linear pattern: see MH 169 (Figure 13: MH169) 4<br />

c18 basic linear pattern: see Z.61.654 (Figure 12) 162 sans zone marker 16<br />

c20 complex linear pattern: see sherd MH360 (Figure 59) similar 163 1<br />

c21 complex linear pattern: see GE 155 (Figure 60) 1<br />

c23 complex linear pattern: see Z.49.751 (Figure 61) 1<br />

c24 complex linear pattern: A11 (Figure 59) 1<br />

c25 complex linear pattern: similar to ch5, see A.2 (Figure 62) 1<br />

cf (crow’s foot) see P205 (Figure 53) compare 197-223 1<br />

ch3 complex linear pattern: see R37.11A.16 (Figure 62) 3<br />

ch5 complex linear pattern: see R37.2A.14 (Figure 64) 157-159 sans zone marker, 307 4<br />

ch6 complex linear pattern: GW.258 (Figure 59) 2<br />

ch7 complex linear pattern: incomplete, see HG.17.8 (Figure 63) 2<br />

ch9 complex linear pattern: see HV.2.124 (?), HV.2.464 (Figure 7), HV.4.175 (Figure 49) 169161 3<br />

ch? complex linear pattern: in triangular pattern category but too small to determine 32<br />

chd complex linear pattern: see MH.185 (Figure 61) 1<br />

chv complex linear pattern: see Z.60.326 (Figure 65) and MH.14 (Figure 66) 187, 188, 190 sans zone markers 45<br />

chz complex linear pattern: see HG.99.061 (Figure 59) 1<br />

cl1 complex curvilinear pattern: R37.7a.21 (Figure 67) 1<br />

cl3 complex curvilinear pattern: see 3711a22 (Figure 67) compare 259 1<br />

cl4 complex curvilinear pattern: see HV.5.214 (Figure 7) 259 with zone marker added? 1<br />

cl5 complex curvilinear pattern: see HV.04.379 (Figure 9) compare 133, 141, 494, 496 1<br />

cl6 complex curvilinear pattern: see GW.159 (Figure 67) 1<br />

cl7 complex curvilinear pattern: see sherd C.14 (Figure 67) 1<br />

cl? curvilinear pattern fragments, not illustrated 3<br />

crw complex linear pattern: see C.101 (Figure 68), post-depositional damage? 1<br />

ct3 (cross-hatch 3) see MH.290 (?), R37.9A.14 (Figure 62), MH.69 (Figure 69) see 230, 237, 7<br />

ctw (cross-hatch/wavy) see MH.290 (Figure 13) 1<br />

cvh (chevron/hatch) see MH.273 (Figure 70) 2<br />

d pattern unknown, element oriented with long axis diagonally oriented to CVA 4<br />

gm3 complex linear geometric pattern: see HV.2.313 (Figure 8) compare 278, 279, 280, 283, 325, 441-448 1<br />

gm4 complex linear geometric pattern: see HV.1.369 (Figure 7) 1<br />

gm5 complex linear geometric pattern: see HG.11.72 (Figure 71) 4<br />

gm6 complex linear pattern: see A.23 (Figure 60) 1<br />

gm7 complex linear pattern: ref R37.7a..13 in data files, faint thin lines, not illustrated 2<br />

gm8 complex linear pattern: See NR.34 (Figure 45) 1<br />

l lateral element, discontinuous, e.g. striation or brush mark 118<br />

lbi transitional between bpt and bpi (one record only) 1<br />

mb multiple stacked continuous latitudinal bands (e.g. Figure 26: HG.21.2) 29<br />

oci discontinuous (widely spaced?) element placed inner edge <strong>of</strong> part 41<br />

oco discontinuous (widely spaced?) element placed outer edge <strong>of</strong> part 2<br />

olb vertical trails <strong>of</strong> paired single impressions (1 example) 1<br />

pc1 complex linear pattern:(?pictorial incised ) see MH44 (Figure 61) 2<br />

pc2 complex linear pattern: see A.24, (Figure 61)<br />

pwc complex linear pattern: see GW055 (Figure 39) 2<br />

rbr complex linear pattern: see R37.11A.2 (Figure 62) 2<br />

rl1 see HV.5.210 (Figure 7) 1<br />

sfl complex linear pattern: see HV.04.185 (Figure 72) 1<br />

wav horizontal wave (created by deformation or impression) (see Figure 42) 96


Figure 51: A band <strong>of</strong> fingernail impressions (opposed pinch), each <strong>of</strong> which is oriented<br />

diagonal to the CVA rather than vertically, or parallel to the CVA.<br />

Figure 52: Deformation, perforation, and the “bnd” incomplete pattern example.<br />

Figure 53: “cf” (Crow’s foot) pattern on the vessel rim above a band <strong>of</strong> pinching.<br />

230


Figure 54: Examples <strong>of</strong> lip impression in pattern “bpi” (band parallel inner-edge <strong>of</strong> lip).<br />

Figure 55: Examples <strong>of</strong> lip impressions laid out in pattern “bpo” (band parallel outeredge).<br />

231


Figure 56: Fragmented examples with lip impressions assigned to “band parallel outer”<br />

pattern.<br />

Figure 57: Examples <strong>of</strong> lip impressions/incision laid out in patterns “bot” (band<br />

opposing top), “bdt” (band diagonal top) and “bpt” (band parallel top), which were<br />

regarded as equivalent in analysis due to non-exclusive nature <strong>of</strong> these descriptions.<br />

232


Figure 58: Pattern expressed using linear arrangements <strong>of</strong> fingernail pinching.<br />

Figure 59: Miscellaneous incised patterns: MH360 is middle row, left-hand column.<br />

GW258 is bottom-right.<br />

233


Figure 60: Examples <strong>of</strong> applied nubbins and some bounded incised patterns (MH259 is<br />

an unbounded pattern).<br />

Figure 61: Decorated sherds with quartz-calcite hybrid granitic temper.<br />

234


Figure 62: Unbounded incised patterns.<br />

Figure 63: Thin incised rims from Hoghoi.<br />

Figure 64: Unbounded incised pattern on the shoulder from Paniavile.<br />

235


Figure 65: An example <strong>of</strong> “chv” pattern, a band <strong>of</strong> unbounded linear triangles filled by<br />

alternating fields <strong>of</strong> parallel lines.<br />

Figure 66: Example <strong>of</strong> “chv” pattern on tall rim (pinching at neck).<br />

236


Figure 67 Curvilinear incised patterns.<br />

Figure 68: Crow’s foot mark, probably post-deposition scratching.<br />

Figure 69: Example <strong>of</strong> cross-hatch pattern “ct3”.<br />

237


Figure 70: Cross hatch pattern “cvh”.<br />

Figure 71: Example <strong>of</strong> pattern “gm5".<br />

Figure 72: Sole example <strong>of</strong> pattern “rl1", a double-line arrangement <strong>of</strong> single repeated<br />

fingernail impressions.<br />

238


For complex patterns expressed graphically, pattern repeat distance was the distance<br />

latitudinally between adjacent homologous points, where the sherd was sufficiently large<br />

to measure this. Pattern repeat distance was not used analytically and is included in case<br />

somebody is interested in this property <strong>of</strong> the assemblages in the future. <strong>The</strong> number <strong>of</strong><br />

vertical repeats <strong>of</strong> patterns comprising stacked multiple bands was recorded along with<br />

miscellaneous other information in the “Notes_1" field <strong>of</strong> “Decoration.db”, and could<br />

easily be systematized into a separate field should anyone so wish in future.<br />

Decorative Elements:<br />

<strong>The</strong> “Element” field specifies that which is repeated in the formulaic patterns: for example,<br />

in the pattern “bp” (lateral band parallel) the pattern consists <strong>of</strong> repeated elements oriented<br />

parallel to each other and perpendicular to the horizon <strong>of</strong> the band, and might be an<br />

opposed-pinch fingernail impression (“opc”), or a deformation <strong>of</strong> the lip into a horizontal<br />

wave (“wav”), or a single impression using a tool or fingernail (“i”). <strong>The</strong> element<br />

represents a single manual decorative operation on the part <strong>of</strong> the potter, each pinch <strong>of</strong> the<br />

fingernails, each impression <strong>of</strong> the tool, each deformation <strong>of</strong> the lip, repeated according to<br />

the pattern. <strong>The</strong> element repeat distance (ERD) specifies the average horizontal spacing,<br />

in mm, between element centres as calculated from two or three measurements, using<br />

calipers, on the sherd. <strong>The</strong>se data were not used in the current analysis, other than for lip<br />

impressions but provide a level <strong>of</strong> descriptive detail that might be <strong>of</strong> use to future analyses,<br />

so were included. Element codes and their meanings are given in Figure 72.<br />

239


Table 16: Decorative elements.<br />

Eleme<br />

nt code<br />

Description count in table<br />

“Decoration”<br />

No element identified (eg surface treatment records) 218<br />

as applied "sausages" or elongate fillets <strong>of</strong> clay 26<br />

ds dentate stamp 7<br />

fi finger/fingertip impression 44<br />

hol hole 82<br />

i linear edge impression not obviously fingernail 141<br />

l linear element (e.g. incision) 239<br />

nub circular nubbin 19<br />

opc opposed pinch fingernail impression 179<br />

pi paired linear impressions 1<br />

psm prismatic or triangular section applied fillet 3<br />

sin single crescentic fingernail impression (not pinched) 27<br />

t applied triangle 1<br />

tfm latitudinal tool forming mark (spatula?) 94<br />

tw? Tool forming marks or striations from wiping? 13<br />

wav wave form, repeated shear-deformation <strong>of</strong> the lip 84<br />

wip wiping or brushing marks 11<br />

ws wavy stamp, probably a shell-edge impression 3<br />

wv3 a closely spaced paired repeat <strong>of</strong> “fi” deformation <strong>of</strong> the lip 8<br />

<strong>The</strong> fields “Mark Length”, “Mark Width” and “Mark Depth” were used to record<br />

dimensions <strong>of</strong> the decorative elements, with maximum depth recorded perpendicular to the<br />

surface <strong>of</strong> the sherd, using a sharpened depth-gauge, in mm, as a positive value regardless<br />

<strong>of</strong> whether decorative element was raised relief, as in the case <strong>of</strong> applied decoration, or<br />

sunken, as in the case <strong>of</strong> incising or punctation.<br />

Measurement conventions for these variables for each element are as given below:<br />

240


Applied fillets (when complete) were measured across their longest and shortest<br />

dimensions for length and width, with depth measured perpendicular to the sherd exterior<br />

surface at a point <strong>of</strong> average depth. dentate-stamp overall length, average width and<br />

average depth were recorded where measurable. Tooth sizes were not recorded but scale<br />

drawings <strong>of</strong> all dentate-stamping are provided (the seven records for dentate-stamping in<br />

the decoration table arise from four sherds with confirmed dentate-stamping and two<br />

sherds with possible dentate-stamping: one sherd yielded two zones <strong>of</strong> dentate-making a<br />

total <strong>of</strong> seven). Fingertip/finger impressions were measured when present as lip<br />

deformations, with length measured around the lip, width measured from the top <strong>of</strong> the lip<br />

to the lowest point <strong>of</strong> the impression. Depth was measured parallel to the top face <strong>of</strong> the<br />

lip, with the base <strong>of</strong> the depth-gauge resting across the length <strong>of</strong> the impression, and the<br />

gauge located at the deepest point. Length <strong>of</strong> holes was measured in their longest<br />

dimension, and width was measured at right angles to the length measurement; depth was<br />

gauged at the deepest point; several readings were averaged where there was variation.<br />

Edge impressions (proportionately deep v-shaped impressions) were measured as<br />

for applied fillets, except for depth, which was measured as for holes. Linear elements were<br />

measured where length was standard for the pattern (e.g. for pattern c18). Nubbins were<br />

measured as for holes, except that depth was measured as for applied fillets, but at<br />

maximum depth. Opposed pinch fingernail impression was measured for length across from<br />

nail-top to nail-top impression at their widest separation, for width at right angles to the<br />

length measurement, and for depth at the deepest point recorded by the depth gauge,<br />

averaging the readings <strong>of</strong> multiple elements where possible. Paired linear impressions<br />

were measured as for linear impressions. Prism applied fillets were measured as for<br />

applied fillets. Rotated tool excision was measured as for finger impression. Single<br />

fingernail impressions were measured as for linear impressions. <strong>The</strong> single example <strong>of</strong> a<br />

pattern made up <strong>of</strong> repeated applied triangles was measured for length across the widest<br />

241


horizontal dimension, and for width in the vertical dimension, and for depth as for other<br />

applique. Wave elements were measured as for finger impressions, and wavelength was<br />

entered as ERD.<br />

Handedness was recorded in many instances but data will not be presented, other<br />

than in the appended database files on CDROM. <strong>The</strong>se data were interesting though as<br />

these provided many clues to vessel orientation during the application <strong>of</strong> decoration<br />

(assuming the potter to be upright) during the process <strong>of</strong> decoration, and also revealed<br />

some variety within the element category <strong>of</strong> opposed pinch fingernail impression.<br />

Additionally, many <strong>of</strong> the deeper fingernail marks were <strong>of</strong> such small size (width) as to<br />

suggest the decorators were women. <strong>The</strong>re is potential for use <strong>of</strong> variation in the shape <strong>of</strong><br />

fingernail impressions as a proxy for genetic variability <strong>of</strong> the potters.<br />

Mark cross-sections across the width <strong>of</strong> the mark (length on the case <strong>of</strong> OPC<br />

fingernail impression) through the depth measurement point were recorded in the “Mark<br />

Section” field as follows: v-section (v), square-section (s), oblique v-section (o), u-shaped<br />

(u), parallel-sided pierced (p), pierced using a conical tool (x), double-v section(w) (as<br />

though mark was incised with a fingernail and fingertip, or with a frayed tool), with “t”<br />

denoting other unclassified cross-sections (specified in “Notes_2").<br />

Transforming Relational Data into a Flat Table:<br />

A flat table with one record (row) per sherd was created from the relational database using<br />

a subset <strong>of</strong> variables pertaining to sherd size, fabric, decoration and form, to provide<br />

summary information on the structure <strong>of</strong> vessel form and decoration as represented by the<br />

sample <strong>of</strong> sherds. (Table 17 shows the data structure and raw data is given in table Flat.db<br />

appended on CD).<br />

Descriptive data for decoration was summarized as a short list <strong>of</strong> decorative<br />

242


attributes, entered into four location fields (decoration-lip = “Part 1", decoration-rim =<br />

“Part 2", decoration-neck = “Part3", decoration-shoulder = “Part 4"). Carination and body<br />

decoration were so highly correlated with upper vessel decoration that these parts were<br />

omitted from the table to prevent multiplicity <strong>of</strong> fields. Attribute value codes used in all<br />

four location-specific decoration fields were as follows :<br />

1. Complex linear motif Class 1 with linear design zone horizontal markers, typically<br />

double-line zone markers<br />

2. Complex linear motif Class2 without linear design zone horizontal markers.<br />

3. Multiple bands <strong>of</strong> opposed pinch fingernail impression (fni)<br />

4. Lateral band <strong>of</strong> punctation<br />

5. Other unclassified decoration not including wiping, brushing or tool-forming marks<br />

6. Plain<br />

(this is a filter attribute)<br />

7. Horizontal finger deformation/impression, discontinuous<br />

8. Wiping, brushing or tool-forming marks (this is a filter attribute as these records<br />

do not constitute decoration in the sense <strong>of</strong> deliberate patterned execution)<br />

9. Staggered opposed inner and outer bands <strong>of</strong> lip impressions forming a wave<br />

pattern without deformation<br />

10. Horizontal finger deformation <strong>of</strong> the lip into a continuous wave<br />

11. Fingernail impression, single band <strong>of</strong> opposed pinching<br />

12. Band <strong>of</strong> impressions along the top <strong>of</strong> the lip<br />

13. Band <strong>of</strong> impressions along the inner edge <strong>of</strong> the lip<br />

14. Opposing bands <strong>of</strong> impressions along both edges <strong>of</strong> the lip, not staggered to form<br />

a wave pattern<br />

15. Band <strong>of</strong> impressions along the outer edge <strong>of</strong> the lip<br />

16. Lateral band <strong>of</strong> applied nubbins<br />

243


Multiple decorative attributes on single vessel part were discarded from the attribute data<br />

for the flat table, but this only affected a small number <strong>of</strong> unusual cases, and does not affect<br />

overall results. Subsequent analyses made use <strong>of</strong> both this flat data structure and the more<br />

comprehensive attributes/variables in the decoration and form tables, through the use <strong>of</strong><br />

multi-table one-to-one and one-to-many database queries.<br />

Parts <strong>of</strong> the vessel not represented by the sherd were identified in “Flat.db” by a<br />

decorative value <strong>of</strong> 99, thus a lip sherd would be coded with a value <strong>of</strong> 99 for each <strong>of</strong> the<br />

missing rim, neck, and shoulder parts. <strong>The</strong> remaining data fields are a transposition <strong>of</strong> basic<br />

variables from the form and thickness tables, separated by vessel part. This draws<br />

information about co-variation by part together into a single table <strong>of</strong> sherd properties for<br />

analysis <strong>of</strong> the structure <strong>of</strong> decoration and form across the vessel.<br />

Chapter Summary and Conclusions:<br />

<strong>The</strong> preceding sections detail the overall data model <strong>of</strong> the ceramic database, and the<br />

contents <strong>of</strong> each <strong>of</strong> the four tables that make up the primary database. A table <strong>of</strong> the vessel<br />

family membership <strong>of</strong> lip and carination sherds, used to estimate a parent breakage<br />

population, is detailed in the next <strong>chapter</strong>. Attributes values were listed and explained in<br />

a series <strong>of</strong> tables and paragraphs pertaining to each attribute, and methods for obtaining<br />

values for variables requiring measured dimensions or counts <strong>of</strong> repeats were explained.<br />

<strong>The</strong> overall objective in designing this database was to try to capture sufficient descriptive<br />

information to be able to draw a picture <strong>of</strong> the vessel as represented by the sherd from the<br />

information recorded, using these rules. Fine structured decorative detail was recorded<br />

formulaically alongside larger, more complex or incomplete patterns that required<br />

illustration.<br />

244


Table 17: Data structure for flat table <strong>of</strong> summary data <strong>of</strong> sherd properties.<br />

Field Name Type Size Key<br />

Id A (alphanumeric) 7 *<br />

Area N (number)<br />

Weight N<br />

Site S (short integer)<br />

Unit (spatial collection S<br />

Code (fabric A 6<br />

Fabric (temper class) S<br />

Form A 2<br />

Sector (EVE N<br />

TYPE S<br />

Decoration Part1 (Lip) S<br />

Decoration Part2 S<br />

Decoration Part3 S<br />

Decoration Part4 S<br />

Lip Type A 1<br />

Lip Angle N<br />

Rim Type A 1<br />

Rim Angle N<br />

Rim Vertical Curve N<br />

Rim depth N<br />

Neck Type A 1<br />

Neck Angle N<br />

NeckVcurve N<br />

Neck Hcurve N<br />

Shoulder Type A 1<br />

Shoulder Angle N<br />

Shoulder Vcurve N<br />

Shoulder Depth N<br />

Carination Type A 1<br />

Carination Angle N<br />

Carination Vcurve N<br />

CariniationHcurve N<br />

Carination notes A 240<br />

Body Type A 1<br />

Body Vcurve N<br />

Body Hcurve N<br />

Lip Thickness N<br />

Rim thickness N<br />

Neck thickness N<br />

Shoulder thickness N<br />

Body thickness N<br />

245


I would like to think that this approach addressed the problem discussed in previous<br />

<strong>chapter</strong>s <strong>of</strong> separate analytical/descriptive schemes for Lapita and for post-Lapita<br />

decorative design, which create an artificial disjuncture in many regional series. This<br />

scheme has been an attempt to bridge that disjuncture, by having a hierarchy <strong>of</strong> descriptive<br />

detail that can capture the broader complexity <strong>of</strong> larger, more complex (and usually<br />

fragmented) designs as well as simpler formulaic repetition <strong>of</strong> basic elements. <strong>The</strong> finer<br />

details <strong>of</strong> technique can be addressed also, down to the physical characteristics <strong>of</strong> potter’s<br />

fingernails, and microscopic differences in technique, measurable by gauge rather than by<br />

eye. This holistic approach to ceramic variation allows the analysis <strong>of</strong> Oceanic pottery to<br />

a level <strong>of</strong> fine detail where variability in potting behaviour from the past can be read<br />

accurately. <strong>The</strong> key, or a major aid at least, to recording at this level <strong>of</strong> detail is a relational<br />

database, where many data themes can be constructed about each part <strong>of</strong> the sherd, while<br />

maintaining the overall relationship <strong>of</strong> these parts to each other and to the whole, without<br />

requiring enormous forms, spreadsheets or tables with hundreds <strong>of</strong> fields <strong>of</strong> information,<br />

most <strong>of</strong> them with missing values due to vessel fragmentation into sherds.<br />

246


CHAPTER 5:<br />

SAMPLE EVALUATION AND QUANTIFYING<br />

Introduction:<br />

SAMPLE SIZE.<br />

As outlined in Chapter 1 the development <strong>of</strong> sherds-as-vessels approaches to ceramic<br />

analysis is recognition <strong>of</strong> the difficulty <strong>of</strong> quantifying sherd samples for comparison unless<br />

they have the same levels <strong>of</strong> brokenness and completeness. Estimation <strong>of</strong> a breakage<br />

population provides a solution to this problem (Felgate & Bickler n.d). Evaluation <strong>of</strong><br />

ceramic sample size and sample representativeness in general continues to be problematic<br />

in Pacific archaeology, where the measured or quantified properties <strong>of</strong> recovered samples<br />

are <strong>of</strong>ten regarded as representative in an unspecified way <strong>of</strong> the properties <strong>of</strong> people’s<br />

lives in the past. In this <strong>chapter</strong> and in subsequent taphonomic and wave-exposure <strong>chapter</strong>s<br />

I try to develop ways to be more specific about the nature <strong>of</strong> that representation, allowing<br />

a more robust behavioral inference from the recovered sherd samples. Objectives in<br />

estimating a breakage population for the Roviana pottery sample are as follows:<br />

• evaluate sample size and significance<br />

• accumulations research: investigate implications for the intensity/duration <strong>of</strong><br />

occupation using the size <strong>of</strong> the estimated breakage population rather than the size<br />

<strong>of</strong> the recovered sample<br />

• estimate the extent <strong>of</strong> post-depositional sherd/pottery attrition between the<br />

breakage population and the extant deposit, as an indication <strong>of</strong> the type/extent <strong>of</strong><br />

sherd-taphonomic process affecting the sites from which the samples were<br />

recovered<br />

247


• characterize the state <strong>of</strong> preservation <strong>of</strong> the deposits (the conditions <strong>of</strong> wave<br />

exposure which are thought to be a key taphonomic factor are explored in Chapter<br />

6). If most <strong>of</strong> the breakage population has vanished even in the sheltered waters<br />

and relative stability <strong>of</strong> sea levels <strong>of</strong> Roviana Lagoon, what are the implications for<br />

the archaeology <strong>of</strong> Lapita in Near Oceania?<br />

• choose a unit <strong>of</strong> quantification: what does the state <strong>of</strong> vessel<br />

brokenness/completeness tell us about how ceramic attributes <strong>of</strong> these sherd<br />

samples should be quantified? If completeness can be shown to be extremely low,<br />

then sherds are independent observations, and sherd counts are likely to be biased<br />

only by differential survival <strong>of</strong> vessel types/parts <strong>of</strong> types, rather than by differences<br />

in brokenness or bias arising from MNI-based counts.<br />

Methods for Establishing Vessel Brokenness and Completeness:<br />

Two characteristics <strong>of</strong> the archaeological sample form the basis <strong>of</strong> an estimate <strong>of</strong> the<br />

breakage population: mean sherd size (measured in EVEs as explained in Chapter 4) as a<br />

measure <strong>of</strong> brokenness (fragmentation); and vessel completeness (Orton et al. 1993:167-<br />

168 and 178). <strong>The</strong> key methodological problem for archaeologists, regardless <strong>of</strong> which<br />

estimation statistic is adopted, is that <strong>of</strong> identifying vessel classes/species/sherd families<br />

in the potsherd sample. Measurement <strong>of</strong> these properties <strong>of</strong> sherd samples has generated<br />

a substantial methodological literature (see for examples Orton 1993, Orton et al.<br />

1993:172, Schiffer 1995c:183-186, Smith 1983:77-79). <strong>The</strong> task <strong>of</strong> identifying sherd<br />

families can be complex, and the prospects <strong>of</strong> a reliable solution being obtained to this<br />

problem vary from sample to sample. This is because pottery is found in an infinite variety<br />

<strong>of</strong> states <strong>of</strong> brokenness and is manufactured with varying degree <strong>of</strong> standardization.<br />

<strong>The</strong> following assumptions were made about the sample:<br />

• <strong>The</strong> sample assemblage is a random sample <strong>of</strong> vessels drawn from a breakage<br />

248


population (this is unlikely to be true, but a good idea <strong>of</strong> the departure from<br />

randomness is obtained in subsequent <strong>chapter</strong>s, and the consequences for this<br />

analysis are discussed)<br />

• Vessel brokenness and vessel completeness can be accurately measured for the<br />

sample.<br />

<strong>The</strong> vessel-randomness <strong>of</strong> a sample and the validity <strong>of</strong> ones measures <strong>of</strong> vessel brokenness<br />

and completeness are central to the level <strong>of</strong> confidence with which the breakage population<br />

estimate can be accepted. It is not a random sample <strong>of</strong> sherd size that is needed.<br />

Taphonomic processes and collection intensity are likely to have structured sherd size, and<br />

as previously discussed, we do not need to concern ourselves with what size that sherds<br />

we do not have were or are. Rather, we must either assume or have reason to believe that<br />

the sherds in the sample assemblage are a random sample <strong>of</strong> vessel membership <strong>of</strong> the<br />

breakage population we wish to estimate.<br />

Collection <strong>of</strong> sherds from broad areas rather than from testpits suggests that the<br />

measured completeness <strong>of</strong> sherd families is unlikely to have been biased by sampling a non-<br />

representative locality in the sites. While test excavation indicated that some pottery<br />

remained buried in some areas <strong>of</strong> some sites, there was no suggestion <strong>of</strong> a buried well-<br />

preserved deposit in any <strong>of</strong> these (Reeve 1989 reports the Paniavile site as such a deposit<br />

but this was not the case by 1996). Turbation by waves and bio turbation are thought to<br />

have created a complex history <strong>of</strong> vertical migration <strong>of</strong> sherds within sediments and some<br />

horizontal movement across the sites, although the lack <strong>of</strong> rolling damage on larger<br />

recovered sherds from the deeper margins <strong>of</strong> sites suggests such horizontal movement was<br />

infrequent.<br />

Lip/rim-based sherd families mostly comprised singleton lip/rim sherds, so spatial<br />

dispersion data could not be obtained from the distribution <strong>of</strong> sherd families.<br />

249


No significant departures from latitudinal circularity were observed among vessel<br />

rims, beyond those attributed to inaccuracy <strong>of</strong> manufacture. Brokenness was measured by<br />

estimating the percentage <strong>of</strong> the circular vessel lip/rim represented by sherds. Lip/rim EVE<br />

was estimated for sherds by curve fitting, using a rim diameter percentage chart marked<br />

in 5% sector intervals, with concentric circles ranging from 20mm radius to 300mm radius,<br />

at intervals <strong>of</strong> 20mm. Where the sherd was too small to estimate lip horizontal radius with<br />

confidence, EVE was estimated using a default radius <strong>of</strong> 140mm, approximating the<br />

average horizontal lip radius recorded for the larger sherds, about which there was little<br />

variation. Sherd sizes ranged up to a maximum lip/rim EVE <strong>of</strong> 25% <strong>of</strong> the vessel<br />

circumference, a recorded minimum <strong>of</strong> 1%, and a recorded average lip and rim EVE for<br />

all sites <strong>of</strong> 5.78% (n=471). Mean EVE by site is shown in Table 18.<br />

Grouping lip/rim sherds into sherd families was the most subjective step in<br />

constructing an estimation <strong>of</strong> the number <strong>of</strong> vessels in a breakage population. Sherds<br />

Table 18: Variation in lip brokenness between sites.<br />

Site Lip sherd count Average EVE (%)<br />

Paniavile 76 4.56<br />

Hoghoi 82 5.16<br />

Miho 60 4.41<br />

Honiavasa 104 7.14<br />

Gharanga 38 5.78<br />

Nusa Roviana 11 7.09<br />

Kopo 3 8.66<br />

Zangana 76 5.93<br />

sorted by site, then sorted further by fabric and by form variation, formed the basis <strong>of</strong> the<br />

primary vessel groupings, which were further subdivided by decoration information. This<br />

required some idea <strong>of</strong> the extent to which fabric and form might vary within a single<br />

250


vessel, and the best information pertaining to these problems was obtained from larger<br />

sherds. Fabric and form attributes were examined at multiple points on such sherds to<br />

develop an assessment <strong>of</strong> the amount <strong>of</strong> within-vessel variation.<br />

No instances were found in which temper <strong>of</strong> different petrographic classes (see<br />

Chapter 4) were incorporated into a single vessel. Grid counts <strong>of</strong> temper density and<br />

mineralogy, under low-powered reflected-light magnification, at multiple points on single<br />

sherds, indicated that there was only slight variation in temper mineralogy (in the relative<br />

proportions <strong>of</strong> mineral grains for example) within vessels, but that temper density could<br />

vary more dramatically within vessels. In sorting lip sherds by fabric, coarse variations in<br />

the relative proportions <strong>of</strong> minerals were used. Colour variation was disregarded, as<br />

surface colour appeared to have been affected by the aerobic regime <strong>of</strong> specific<br />

depositional context, and colour zonation in the break could be expected to vary within a<br />

single vessel as a result <strong>of</strong> firing variability.<br />

Form variation <strong>of</strong> the lip and rim within vessels was examined qualitatively by<br />

looking at larger sherds where a greater proportion <strong>of</strong> the vessel was represented (up to<br />

about 30%). Slight variation in form around the lip/rim circumference were noted, and<br />

small variations <strong>of</strong> this order <strong>of</strong> magnitude were ignored when sorting fabric groups into<br />

vessel groups based on form.<br />

Sherds for which the edges <strong>of</strong> the lip were weathered, making decorative analysis<br />

impossible, were generally small and water-rolled, and had been collected along the<br />

strandline. <strong>The</strong>se were omitted from the analytical sub-sample used to estimate<br />

completeness.<br />

It was noted that few if any <strong>of</strong> the more decorated lip/rim sherds could be matched<br />

to any other, suggesting either that decoration changed around the circumference <strong>of</strong> the<br />

vessel lip/rim, or that vessel completeness for all site assemblages was very low. Some<br />

sherds showed latitudinal decorative discontinuities <strong>of</strong> this type, but these discontinuities<br />

were restricted to a small range <strong>of</strong> incised motifs, omitted from subsequent analyses.<br />

251


This relationship between the amount <strong>of</strong> decorative information on lip and rim<br />

sherds pertaining to vessel membership, and the sherd count for vessel groups, was<br />

systematized by counting the number <strong>of</strong> latitudinal bands <strong>of</strong> repeated decorative elements<br />

that made up the total decoration <strong>of</strong> the lip and rim. <strong>The</strong>se were typically bands <strong>of</strong> repeated<br />

parallel edge-impression along the edges/top <strong>of</strong> the lip, or bands <strong>of</strong> punctation or incised<br />

decoration around the rim. <strong>The</strong> properties <strong>of</strong> the tool used to make the decorative marks<br />

and the average spacing and depth <strong>of</strong> marks were also used to differentiate between vessels<br />

in some cases. Here again, this depended on assessing through examination <strong>of</strong> larger<br />

sherds, the amount <strong>of</strong> variation one might expect in this regard within vessels. For all sites,<br />

this system yielded the data in Table 19 (vessel membership data is appended on CD in<br />

“Vessels.db”).<br />

Table 19: Selection <strong>of</strong> sample for estimating vessel completeness.<br />

Sherd type/<br />

# bands<br />

decoration<br />

Sherd<br />

count<br />

Vessel count Min. # per<br />

vessel<br />

252<br />

Max. # per<br />

vessel<br />

Mean # pieces<br />

per vessel<br />

(completeness<br />

lip/rim (0 bands) 91 53 1 14 1.72<br />

lip/rim (1 band) 264 166 1 15 1.58<br />

lip/rim (2 bands) 68 65 1 2 1.05<br />

lip/rim (3 or more<br />

bands)<br />

lip/rim (2 or more<br />

bands)<br />

28 25 1 2 1.12<br />

96 89-90 1 2 1.07-1.08<br />

Undecorated lips or lips with a single band <strong>of</strong> decoration (usually parallel edge<br />

impressions) yielded up to 15 pieces per vessel. <strong>The</strong> sample <strong>of</strong> more decorated sherds by<br />

contrast yielded lower vessel completeness, with a maximum <strong>of</strong> two lip/rim sherds per<br />

vessel, and an average <strong>of</strong> 1.07-1.08 sherds per vessel, despite similar brokenness. This<br />

lower level <strong>of</strong> completeness is considered to be a more accurate reflection <strong>of</strong> the state <strong>of</strong><br />

the assemblages than the figures for the plain or slightly decorated lips. <strong>The</strong> large vessel<br />

groups for the plain or slightly decorated lips are inferred to be faulty attributions <strong>of</strong>


sherds to vessel groups as a result <strong>of</strong> inadequate information with which to construct<br />

groups. This inference depends on the assumption that stylistic variation between sherds<br />

results from idiosyncratic recombination <strong>of</strong> the potters' repertoire <strong>of</strong> continuous horizontal<br />

bands <strong>of</strong> decoration for each pot constructed, and also on the related assumption that<br />

differences between sherds do not result from changes in decoration around the<br />

circumference <strong>of</strong> the vessel.<br />

Simulation Approach:<br />

<strong>The</strong> number <strong>of</strong> pieces into which a population <strong>of</strong> vessels breaks is variable. <strong>The</strong><br />

probabilities <strong>of</strong> individual vessels being represented in an archaeological sample are thus<br />

unequal. For pottery, the “death” <strong>of</strong> a pot (breakage) generally implies at least two pieces,<br />

although there are exceptions. Moreover, breakage is not necessarily a single event. As<br />

Orton states,<br />

“It is obvious that sherds can either stay the same size or become smaller<br />

(through breakage) over time (Orton et al. 1993:176).”<br />

In the latter case, these would not only become more numerous (increasing probability <strong>of</strong><br />

survival/capture in the archaeologist’s sample), but once below the detection size<br />

threshold, would become less likely to be found (decreasing probability <strong>of</strong> capture in the<br />

archaeologist’s sample). This last factor constrains the minimum size <strong>of</strong> sherds commonly<br />

recovered by archaeologists. An additional size-related factor in survival/detection is that<br />

smaller fragments are potentially more mobile, either within sediments as a result <strong>of</strong><br />

turbation or on the sediment surface due to fluvial or aeolian transport.<br />

While it is possible to envisage either progressive breakage (see Orton 1982 for<br />

a simulation approach to progressive breakage) or a state <strong>of</strong> breakage stasis, it is not<br />

possible to know from the sample assemblage which <strong>of</strong> these situations applied to a<br />

pottery breakage population; these are equifinal in that either progressive breakage or<br />

253


static breakage could produce any found broken assemblage. <strong>The</strong> state <strong>of</strong> brokenness in<br />

which a pottery breakage population existed is usually unknowable from the material<br />

record because the creation <strong>of</strong> a pottery breakage population is usually a diachronic<br />

process, and archaeologists only have access to the deposited/fossil assemblage through<br />

their sample assemblage as a synchronic (or nearly so) observation. A vast array <strong>of</strong> small<br />

breakage and re-breakage events may have occurred (or not) over the period prior to<br />

sample recovery, except in exceptional circumstances, such as where people or natural<br />

formation processes have curated or preserved pots in their death-state upon breakage,<br />

either acting continuously in this way over time, or where by some instantaneous process<br />

a whole assemblage <strong>of</strong> vessels is broken at once and lain undisturbed since (as at Pompeii).<br />

In such circumstances, preservation would be so good that there would be no need for<br />

estimation approaches.<br />

It is the property <strong>of</strong> potsherd samples that whole objects are broken into a number<br />

<strong>of</strong> fragments that allows us to establish vessel representation fraction (the fraction <strong>of</strong> the<br />

breakage population represented in the sample). If an assemblage <strong>of</strong> unbroken pots were<br />

recovered we would have no information on whether others existed at that location in the<br />

past and were carried <strong>of</strong>f in the intervening years. <strong>The</strong> more broken an assemblage, the<br />

more the survival chances <strong>of</strong> each vessel as a fragment improve (to a degree) as do the<br />

chances that some fragment <strong>of</strong> any individual vessel will make it into the archaeologist's<br />

sample.<br />

To reiterate, basic assumptions are:<br />

1: <strong>The</strong> sample assemblage is a random sample <strong>of</strong> vessel membership <strong>of</strong> the<br />

breakage population,<br />

2: vessel brokenness and completeness can be accurately measured from the sherd<br />

sample.<br />

Sample brokenness was defined as the mean number <strong>of</strong> rim sherds per vessel, and was<br />

254


calculated by dividing 100% by the mean EVE <strong>of</strong> sherds in the sample assemblage. As<br />

EVE is an estimate, assemblage brokenness is also an estimate, although no attempt is<br />

made here to calculate precision <strong>of</strong> that estimate. Vessel completeness was calculated as<br />

the estimated mean number <strong>of</strong> pieces per vessel in the archaeological sample, or the<br />

calculated mean number <strong>of</strong> pieces per vessel in any simulated pottery sample<br />

<strong>The</strong> simulation sherd sampling fraction was defined as the number <strong>of</strong> sherds one<br />

would have to remove from a simulated breakage population <strong>of</strong> complete vessels<br />

fragmented as observed in the sample assemblage, to achieve the level <strong>of</strong> completeness<br />

observed. This definition implies that brokenness <strong>of</strong> the assemblage was static in the past<br />

rather than progressive, and means that rather than asking the question 'what fraction <strong>of</strong><br />

an actual population <strong>of</strong> sherds near the time <strong>of</strong> breakage <strong>of</strong> the pottery is present in the<br />

sample?', the question posed in our approach is rather 'If the pottery in this deposit had<br />

existed since initial breakage in the state <strong>of</strong> brokenness observed in the sample, what<br />

fraction <strong>of</strong> the initial sherd population would the sample represent?'<br />

<strong>The</strong> simulation vessel representation fraction is the fraction <strong>of</strong> the simulated<br />

breakage vessel population represented at any given simulated sherd sampling fraction.<br />

While sherd sampling fraction is a stipulation device permitting the calculation <strong>of</strong> vessel<br />

representation in samples, based on the expected number <strong>of</strong> similar-sized pieces in the<br />

whole breakage population, vessel representation fraction seems to be a valid measure <strong>of</strong><br />

the relationship between sample and parent population, and informs as to the fraction <strong>of</strong><br />

the breakage population represented by an archaeological vessel/sherd sample. <strong>The</strong> sherd<br />

sampling fraction range is the range <strong>of</strong> outcomes in the simulation for which vessel<br />

completeness is consistent with the archaeological sample. <strong>The</strong> vessel representation<br />

fraction range is the range <strong>of</strong> simulation vessel representation fractions that are consistent<br />

with the vessel completeness <strong>of</strong> the archaeological sample.<br />

In the simulation, brokenness is treated as static, and is determined by specifying<br />

255


the number <strong>of</strong> vessels and the number <strong>of</strong> sherds in a simulated breakage population. Vessel<br />

completeness in these simulated breakage populations will <strong>of</strong> course always be 100%.<br />

Brokenness is made consistent with that observed in the archaeological sample. <strong>The</strong> goal<br />

<strong>of</strong> the simulation is to establish, for Roviana sample assemblage brokenness, the shapes <strong>of</strong><br />

the curves and the range <strong>of</strong> likely values that relate:<br />

a) simulation sherd sampling fraction to vessel representation fraction, and<br />

b) simulation vessel completeness to simulation sherd sampling fraction.<br />

This is achieved by simulating a large number <strong>of</strong> sherd sample assemblages with a level <strong>of</strong><br />

brokenness consistent with the archaeological sample. Vessel membership <strong>of</strong> sherds in all<br />

samples is assigned randomly, but overall, is consistent with the expected number <strong>of</strong> pieces<br />

per vessel at that sampling fraction. Completeness (mean number <strong>of</strong> sherds per vessel) and<br />

vessel representation fraction are calculated for each sample and are displayed in a plot to<br />

show the relationship between vessel completeness and vessel representation fraction. By<br />

this means a good estimate can be gained <strong>of</strong> the sherd and vessel sampling fractions in the<br />

simulation consistent with the vessel completeness and brokenness estimated for the<br />

archaeological sample.<br />

To reiterate the statement made earlier, this is not a suggestion that the breakage<br />

population had the particular static level <strong>of</strong> brokenness <strong>of</strong> the archaeological sample, and<br />

<strong>of</strong> all the simulated samples. <strong>The</strong> question <strong>of</strong> how many bits someone's pottery broke into<br />

when they dropped it is probably <strong>of</strong> little historical or archaeological interest, and is<br />

usually unknowable. Stipulating a level <strong>of</strong> brokenness consistent with the archaeological<br />

sample assemblages is rather a way <strong>of</strong> calculating, which allows the question 'how much<br />

needs to be taken away from the whole to give it the characteristics <strong>of</strong> the sample? We<br />

also do not really have to be concerned with what size any extant missing sherds are or<br />

were in the total discard assemblage, as ultimately it is vessel representation fraction we<br />

are after as an entry to quantification <strong>of</strong> the breakage population, rather than sherd<br />

256


sampling fraction itself.<br />

Estimating the number <strong>of</strong> vessels in a breakage population requires approaches that<br />

allow a variable number <strong>of</strong> members in classes (for example Chao 1984, Chao & Lee<br />

1992). Accordingly, the simulated samples are created as datasets containing a variable<br />

number <strong>of</strong> sherds per vessel. In the simulated breakage population, sherds are randomly<br />

assigned to vessel membership based on a probability derived from brokenness as measured<br />

by the archaeological sample mean EVE.<br />

Statistical Approaches:<br />

While the simulation approach is useful for elucidating the relationships between<br />

brokenness, completeness, and vessel representation in the sample, a number <strong>of</strong> statistical<br />

algorithms apply to the problem <strong>of</strong> estimating a breakage population from a sherd sample<br />

(estimating the number <strong>of</strong> classes/species in a population). <strong>The</strong>se have the advantage <strong>of</strong><br />

producing formal confidence limits. Of particular relevance to the problem are the capture-<br />

recapture statistics used in ecological research where a sample <strong>of</strong> animals are captured,<br />

tagged and then released back into the population. Once they have redistributed across the<br />

landscape, animals are captured and then released again in a series <strong>of</strong> trials. <strong>The</strong> frequency<br />

<strong>of</strong> capture <strong>of</strong> particular tagged animals is used to generate an estimate <strong>of</strong> the total<br />

population. While there is no attempt here to summarise the vast literature on this topic<br />

and the complexities involved, the following points are particularly relevant to the<br />

archaeological problem.<br />

<strong>The</strong> capture-recapture process works because the probability <strong>of</strong> capturing a tagged<br />

animal from a population tells us something about the population size. Getting one tagged<br />

animal from a single sample does not really tell you anything about an unknown<br />

population size. However, the more tagged animals you get and the more times the same<br />

257


tagged animal is recaptured in successive trials, the more certain you can be about the size<br />

<strong>of</strong> population that it probably derives from. <strong>The</strong>se statistics have been reworked to<br />

estimate the number <strong>of</strong> species in a population, in which form these are directly applicable<br />

to the problem <strong>of</strong> estimating the number <strong>of</strong> pottery vessels represented by a sherd sample.<br />

Here the number <strong>of</strong> ecological capture trials is analogous to the maximum number <strong>of</strong><br />

pieces from a vessel in the archaeologist's sample, with incidence <strong>of</strong> singleton sherds<br />

comprising the sample <strong>of</strong> animal captured in the first trial, the incidence <strong>of</strong> doubletons<br />

(pots represented by two sherds in the recovered sample) being the second capture sample,<br />

the incidence <strong>of</strong> tripletons (pots represented by three sherds) the third, and so on.<br />

This approach does not make explicit use <strong>of</strong> the sherd-size information inherent in<br />

pottery samples, which informs on the probability <strong>of</strong> capture (brokenness), which seems<br />

wasteful <strong>of</strong> information. Such information on the probability <strong>of</strong> capture could potentially<br />

be used to narrow the confidence limits obtainable for a given sample size, or to reduce the<br />

sample size required to obtain meaningful confidence limits.<br />

<strong>The</strong> Chao (1984) estimate was tested experimentally (Felgate & Bickler n.d) using<br />

data simulated from hypothetical breakage populations. Capture-recapture data calculated<br />

from these simulated samples was used to estimate (with confidence limits) the known<br />

breakage population, using the program “EstimateS”(Colwell 1997). <strong>The</strong> estimates from<br />

the Chao 1984 statistic were compared with the known breakage population.<br />

<strong>The</strong> key findings were that:<br />

• Sherd sample fractions <strong>of</strong> 5% or more allow a close estimate <strong>of</strong> the breakage<br />

population (this figure is a product <strong>of</strong> the size <strong>of</strong> the breakage population used in<br />

the trial, and could be improved using a larger population)<br />

• Even at 2% sampling fraction, the result is useful if the breakage population is<br />

large<br />

• Higher mean EVE (less broken assemblages) make the statistic less accurate,<br />

258


equiring a larger sampling fraction (but at the opposite extreme, in very broken<br />

assemblages, although the statistic works well, the data will be impossible to<br />

acquire; if the pottery is too broken the measurement <strong>of</strong> EVE itself is<br />

compromised).<br />

In contrast to the simulation method, the Chao statistic does not require discrete<br />

information on assemblage brokenness. <strong>The</strong> structure <strong>of</strong> the capture-recapture data<br />

provides this information. While the Chao statistic appeared to give good results only if<br />

the sherd sampling fraction is greater than 5% <strong>of</strong> the breakage population, this was a<br />

sample size effect. If the simulated breakage populations used by Felgate and Bickler had<br />

been larger than 10,000 sherds, confidence limits ought to have narrowed at lower<br />

sampling fractions.<br />

Simulation Results:<br />

Using the two-or-more-bands-<strong>of</strong>-decoration approach outlined above, completeness was<br />

estimated to be between 1.07 and 1.08 mean sherds per vessel. Brokenness was calculated<br />

by taking the average EVE <strong>of</strong> all lip sherds in the site regardless <strong>of</strong> the amount <strong>of</strong><br />

decoration present (5.78%). Site samples were combined to boost sample size. <strong>The</strong><br />

question being asked <strong>of</strong> the following simulation is therefore “What is the total breakage<br />

population for these six sites?”<br />

An initial simulated breakage assemblage <strong>of</strong> 10 000 sherds forming 578 vessels was<br />

stipulated, consistent with overall sample Brokenness (5.78% EVE). Initial simulation with<br />

500 sampling fractions from 0% to 100% indicated that vessel completeness <strong>of</strong> 1.08 sherds<br />

per vessel corresponds with simulation sherd sampling fraction <strong>of</strong> somewhere less than 3%<br />

(Figure 73).<br />

259


A second, detailed simulation was run using an enlarged breakage population <strong>of</strong><br />

Figure 73 Roviana vessel completeness against sampling fraction, sampling<br />

fraction against vessel representation fraction; random assignment <strong>of</strong> shreds to<br />

vessels; mean EVE <strong>of</strong> 5.78%.<br />

100,000 sherds (5780 vessels) at 11 sampling fractions between 0% and 2%, iterated 100<br />

times, plotting 1100 experiments in total. Vessel completeness <strong>of</strong> 1.07-1.08 sherds per<br />

vessel correspond on the simulation plot with vessel representation fraction <strong>of</strong> between<br />

8.9% and 17.5% (Figure 74).<br />

260


<strong>The</strong> total archaeological rim/lip sample was 471 sherds. Based on vessel completeness<br />

estimated to be from 1.07 to 1.08 sherds per vessel, this sample was inferred to physically<br />

represent between 436 and 440 vessels. This suggests limits for a total breakage population<br />

for the analyzed sites <strong>of</strong> 3259-4944 at 1.07 completeness, or at 1.08 completeness, 2491-<br />

4152 vessels. <strong>The</strong> combined approximate limits <strong>of</strong> the breakage population for this run <strong>of</strong><br />

the simulation program would be 2491-4944 vessels.<br />

Repeating the simulation two more times yielded vessel representation fractions<br />

(using completeness <strong>of</strong> 1.07-1.08) <strong>of</strong> 9.6%-19%, and 8.8%-18.9%. Combining the three<br />

ranges obtained yields vessel representation fraction range <strong>of</strong> 8.8%-19%, suggesting a<br />

more conservative approximation <strong>of</strong> breakage population confidence limits at 2sd would<br />

be 2295-5000 vessels.<br />

For comparison with the Chao calculation given below, a single completeness<br />

Figure 74: Roviana vessel completeness against vessel representation fraction;<br />

random assignment <strong>of</strong> sherds to vessels; mean EVE <strong>of</strong> 5.78%.<br />

261


value <strong>of</strong> 1.08 was used, yielding a total vessel count <strong>of</strong> 436 in the sample <strong>of</strong> 471 rim<br />

sherds. Vessel representation fraction averaged limits were 10.5%-17.5%, 11.2%-19%,<br />

and 10.6%-18.9%. Combined vessel representation fraction limits at this level <strong>of</strong><br />

completeness were 10.5% to 19%. <strong>The</strong> total sample assemblage <strong>of</strong> 436 vessels (at 1.08<br />

completeness) represents a breakage population <strong>of</strong> 2295-4152 vessels.<br />

Statistical Results:<br />

<strong>The</strong> results <strong>of</strong> analysis <strong>of</strong> the same Roviana data using the Chao (1984) statistic (Table 20)<br />

are similar. <strong>The</strong> results do not coincide, but the ranges overlap, which is encouraging,<br />

especially as the simulation results indicate that the Chao statistic is operating at its limits<br />

<strong>of</strong> usefulness at such small sampling fractions. <strong>The</strong> 89 vessels in the sample yielded parent<br />

population limits at 1s.d. <strong>of</strong> 293.38-716.16 vessels, or vessel representation fraction limits<br />

<strong>of</strong> 0.3034-0.1243. (At 2s.d., i.e. ±422, the breakage population represented by the 89<br />

vessels in the sample would be 83-927, illustrating how at these very small sampling<br />

fractions and limited sample sizes the Chao statistic has insufficient information on the<br />

probability <strong>of</strong> “capture” in the sample to give meaningful confidence limits.) <strong>The</strong> 436<br />

vessels in the total rim/lip sherd sample assemblage, at 1s.d., represent a breakage<br />

population <strong>of</strong> 1437-3508 vessels (compare this to a breakage population <strong>of</strong> 2295-4152<br />

vessels using the same data in the simulation).<br />

Table 20: Breakage population estimate for the combined Roviana highly decorated lip<br />

sample using the statistic <strong>of</strong> Chao 1984.<br />

Vessel count Sherd count Singletons Doubletons Breakage population 1s.d.<br />

89 96 82 7 504.77 211.39<br />

262


Discussion:<br />

In cases where the Chao statistic yields estimates with reasonably tight confidence limits,<br />

it is clearly a more systematic solution to the estimation problem than our simulation<br />

approach as it permits an estimate with confidence limits without the archaeologists<br />

specifically inputting brokenness information, which may itself be subject to error. In the<br />

case <strong>of</strong> the Roviana data the sample size is too small for the structure <strong>of</strong> the multiple<br />

capture data to inform as to the probability <strong>of</strong> capture, leading to an overly large standard<br />

deviation (at two standard deviations the 89 vessels in the sample would represent between<br />

83 and 927 vessels, which is not a very convincing result). <strong>The</strong> Chao estimate does rely on<br />

the structure <strong>of</strong> the capture-recapture data to establish the brokenness <strong>of</strong> the sample. This<br />

structure breaks down at small sampling fractions and small sample sizes. In these<br />

circumstances the simulation approach provides a better estimate than the Chao statistic,<br />

as brokenness is independently established from the EVE characteristics <strong>of</strong> the sample <strong>of</strong><br />

lip/rim sherds.<br />

Summary and Conclusions:<br />

Two methods <strong>of</strong> estimating a breakage population were applied, a simulation approach,<br />

and a statistical approach using the algorithm developed by Chao (1984). Both require that<br />

sherd families be identified in the sample. <strong>The</strong> simulation approach requires a measure <strong>of</strong><br />

completeness (in mean number <strong>of</strong> pieces per vessel) and brokenness (expected number<br />

<strong>of</strong> pieces per vessel) from the archaeological sample. <strong>The</strong> simulation approach proceeds<br />

by simulating many samples <strong>of</strong> a breakage population with a degree <strong>of</strong> brokenness<br />

consistent with the archaeological sample, to develop plots <strong>of</strong> the relationship between<br />

vessel completeness and vessel representation fraction. <strong>The</strong> measured vessel completeness<br />

263


from the archaeological sample is then compared with these plots to discover the<br />

corresponding vessel representation fraction indicated by the simulation. Dividing the<br />

number <strong>of</strong> vessels represented in the archaeological sample by this vessel representation<br />

fraction range yields the likely limits <strong>of</strong> vessels in the breakage population.<br />

<strong>The</strong> statistical approach uses a version <strong>of</strong> the capture-recapture approach for<br />

estimation <strong>of</strong> animal populations, adapted to the problem <strong>of</strong> estimating the number <strong>of</strong><br />

species in a population. In this application to archaeological pottery, the number <strong>of</strong> vessel<br />

sherd families in a breakage population is equivalent to the number <strong>of</strong> species in a<br />

population, and the presence <strong>of</strong> multiple sherds from the same vessel in the archaeological<br />

sample is equivalent to multiple capture events. This approach was tested on a range <strong>of</strong><br />

simulated breakage populations and simulated archaeological samples, and for the sample<br />

sizes tested, found to be accurate for sherd sampling fractions <strong>of</strong> 5% or more. Moreover,<br />

the statistic was more accurate when the vessel assemblage was more broken because <strong>of</strong><br />

the enlarged sherd sample for each vessel, which increases the probability <strong>of</strong> multiple<br />

“captures” at any given sherd sampling fraction.<br />

Drawing a theoretical distinction between a breakage population and a discard<br />

assemblage means that there is no required assumption that all <strong>of</strong> the sherds <strong>of</strong> pottery<br />

were discarded at the site. Sherds discarded or destroyed <strong>of</strong>f site are simply sherds that did<br />

not make it into the sample. Only when seeking a history <strong>of</strong> the pottery outside the sample<br />

assemblage does such transport or destruction become significant.<br />

Both estimation approaches require close attention to the nature <strong>of</strong> the total<br />

sampling processes. A vessel-random selection <strong>of</strong> sherds from the breakage population is<br />

the ideal. In archaeological terms this is most likely the case when sherds from the same<br />

vessel show a high degree <strong>of</strong> dispersion across the site. This suggests the approach is most<br />

suited to analysis <strong>of</strong> surface collections, or collections from large areal excavations, where<br />

dispersion information is attainable. Alternatively, other sampling approaches to<br />

264


archaeological sites such as multiple test excavations across a site might yield adequate<br />

information on the dispersion <strong>of</strong> sherd families.<br />

For both approaches, the accuracy <strong>of</strong> estimates <strong>of</strong> breakage population obtained<br />

depends on the accuracy <strong>of</strong> the estimate <strong>of</strong> vessel completeness. For the simulation<br />

approach an estimate <strong>of</strong> sample brokenness is also required. <strong>The</strong> Roviana data illustrate<br />

some <strong>of</strong> the difficulties in attaining reliable data <strong>of</strong> this sort. This requirement places a<br />

heavy burden on the analyst to develop reliable ways <strong>of</strong> sorting a sherd assemblage into<br />

vessel lots. This is not a simple analytical task and there is vast scope for methodological<br />

development in this area.<br />

For the overall Roviana sample, where vessel completeness is uniformly low, we<br />

can infer from the simulation that around 99% <strong>of</strong> the sherd material <strong>of</strong> the breakage<br />

population has not made it into the sample, and secondly, can appreciate that all is not lost,<br />

and that means exist for estimating the size <strong>of</strong> the total breakage population represented<br />

by the sample collected.<br />

Sample sizes were too small to estimate the composition <strong>of</strong> the breakage<br />

population for individual sites, and an argument is made elsewhere that some production<br />

styles could be better estimated from other vessel parts such as necks or carinations, due<br />

to fragile rim form (elaborated in Chapters 8 and 9).<br />

<strong>The</strong> extent <strong>of</strong> sherd attrition shows that assemblage composition is potentially<br />

biased and is most likely comparable only with other assemblages which have been subject<br />

to a similar taphonomic/sampling regime. It is clear that the vast bulk <strong>of</strong> sherdage deriving<br />

from the breakage population has not made it into the archaeologists sample. <strong>The</strong> count<br />

(assuming breakage stasis) or weight <strong>of</strong> sherds recovered would underestimate the<br />

intensity/duration <strong>of</strong> occupation by a factor <strong>of</strong> 100. <strong>The</strong> estimated number <strong>of</strong> vessels<br />

physically represented in the archaeological sample under-represents the size <strong>of</strong> the<br />

breakage population by a factor <strong>of</strong> between five and twelve.<br />

265


While some pottery was found buried in sediments in some sites, through test<br />

excavation, this could account for only a tiny fraction <strong>of</strong> the missing pottery, and as<br />

sample collection <strong>of</strong> the reef-flat surface scatter was intensive, it is inferred that the bulk<br />

<strong>of</strong> the breakage assemblage pottery had been removed by taphonomic processes.<br />

Observation <strong>of</strong> exotic pyroxene minerals in the local calcite sand <strong>of</strong> one site (Hoghoi)<br />

suggests much <strong>of</strong> the breakage population has dis-aggregated entirely into constituent<br />

mineral grains.<br />

<strong>The</strong> potential for bias in the composition <strong>of</strong> recovered sample assemblages is<br />

therefore inferred to be high for the sites as a group. <strong>The</strong> conclusion for any analyses that<br />

use relative abundance <strong>of</strong> styles or attributes (frequency seriation for example) must be<br />

that external comparisons should be restricted to sites that have undergone similar<br />

taphonomic processes. Direct comparisons should be avoided even with sites that seem<br />

to have similar cultural formation processes, but different taphonomic regimes (for<br />

example the Mussau sites).<br />

For accumulations research, the estimate will under represent some fragile lip<br />

forms, and also cannot represent types or sites that have not made it into the sample at all.<br />

It is entirely possible, indeed likely, that many more than 4000 vessels were broken around<br />

the intertidal zone <strong>of</strong> Roviana lagoon in the past, but the result obtained is sufficient to<br />

confirm the intuitive understanding that most <strong>of</strong> the sherdage must have gone, in order to<br />

produce the observed pattern, and that at least five times as many vessels as the number<br />

observed in the sample were broken here.<br />

How fragile is this site type? <strong>The</strong> degree <strong>of</strong> wave exposure for sites is compared<br />

in Chapter 6 but all sites are broadly similar due to a broadly comparable situation within<br />

the Lagoon. It is clear from the simulation that around 99% <strong>of</strong> the physical quantity <strong>of</strong><br />

ceramics has not been captured in the sample, and as sampling was quite intensive, aiming<br />

at total collection <strong>of</strong> lip sherds, it is clear that site preservation even in the sheltered waters<br />

266


<strong>of</strong> the Lagoon is marginal, and that this type <strong>of</strong> site is almost entirely removed by<br />

taphonomic processes.<br />

This result suggests that aceramic lithic scatters in similar locations may have had<br />

a ceramic component in the past, removed by differential weathering <strong>of</strong> ceramics and<br />

lithics. Had similar sites been located in more wave-exposed settings elsewhere in Near<br />

Oceania, we can be certain that few <strong>of</strong> these would have survived to be detectable as<br />

ceramic scatters by an archaeologist today. <strong>The</strong> poor state <strong>of</strong> preservation <strong>of</strong> the Roviana<br />

intertidal-zone sites suggest settlement patterns, site density etc for this site type cannot<br />

be read directly from survey results, and that preservation and visibility factors ( probability<br />

<strong>of</strong> detection) must be carefully considered to read survey results.<br />

Finally, the low state <strong>of</strong> completeness <strong>of</strong> the Roviana vessels indicates that sherd<br />

counts <strong>of</strong> decorative attributes are likely to be independent in the vast majority <strong>of</strong> cases,<br />

and that sherd count is the best unit for counting decorative attributes (quantifying sample<br />

size) for seriation analysis.<br />

267


268


Introduction:<br />

CHAPTER 6:<br />

WAVE EXPOSURE<br />

Systematic consideration <strong>of</strong> the effects <strong>of</strong> wave exposure on archaeological deposits are<br />

rare or superficial in recent Oceanic archaeology, despite the situation <strong>of</strong> the subject in the<br />

largest wave basin on the planet. In the 1920s the prevailing paradigm in Oceanian<br />

archaeology was that there was no Oceanic subsurface archaeology due to the erosional<br />

effects <strong>of</strong> cyclones, floods, high temperatures and damp conditions (an example <strong>of</strong> this<br />

paradigm can be found in Linton’s introduction to his “Archaeology <strong>of</strong> the Marquesas”<br />

published in 1925 (Linton 1925:3-4), but from the 1930s through to the present the<br />

pendulum <strong>of</strong> scientific opinion has swung to the opposite extreme, with the archaeological<br />

record read relatively directly, and minimal attention has been paid to formation processes<br />

in general and the degree to which wave exposure structures Oceanian archaeological<br />

distributions. A sample surveying approach by contrast demands attention to probability<br />

<strong>of</strong> site detection, including preservation and visibility, as lenses through which the<br />

archaeological record is read more accurately.<br />

In Chapters 7 and 11 the ceramic scatters are viewed as lag deposits, evidenced by<br />

a high degree <strong>of</strong> size sorting. <strong>The</strong> value in constructing this model <strong>of</strong> wave exposure in<br />

the current <strong>chapter</strong> lies not so much in being able to test it against sherd size data, but in<br />

having some means other than the sizes <strong>of</strong> recovered sherds to assess relative exposure.<br />

Sherd size is subject to a collection intensity effect, making highly size-sorted<br />

assemblages difficult to distinguish from poorly sorted assemblages unless scrupulous<br />

control for collection intensity can be maintained (difficult unless one knows the entire<br />

269


history <strong>of</strong> collection <strong>of</strong> the site). Modelling wave-exposure allows a cross-check on size<br />

data, and promotes the view that factors additional to natural transforms are structuring<br />

the taphonomic data in assemblages (in the next <strong>chapter</strong> some conflicts between expected<br />

size distribution from wave exposure and actual size distribution <strong>of</strong> the archaeological<br />

sample will be highlighted which illustrate this).<br />

Review:<br />

Some studies in which wave exposure receives particular mention are reviewed below<br />

(passing reference is made in many studies to wave-laid sediments in excavations, but these<br />

are reviewed elsewhere (Tarlton 1996).<br />

Green’s survey <strong>of</strong> the Surville peninsula on Makira (San Cristobal) in 1970 noted<br />

the dramatic effects <strong>of</strong> 1971 cyclone waves on coastal rockshelters. For Ulawa in the<br />

southeast Solomon Islands (Hendren 1976, Ward 1976) Ward notes evidence for exposure<br />

<strong>of</strong> this coast to cyclone waves, and effects on terrestrial site preservation. <strong>The</strong> occupation<br />

sites excavated by Ward postdated about 700AD. Results suggest terrestrial pottery sites<br />

<strong>of</strong> any period are rare or absent from the extant archaeological record.<br />

Lilley in a study <strong>of</strong> islands in the Vitiaz Strait noted the likely impact <strong>of</strong> the 1888<br />

Ritter tsunami on the north coast <strong>of</strong> Umboi, parts <strong>of</strong> coastal Sakar, western New Britain<br />

and the low-lying Siassi Islands, and coasts <strong>of</strong> Malai and Tuam islands (Lilley 1986:20),<br />

and suggested that low levels <strong>of</strong> site visibility and survival could be expected as a result <strong>of</strong><br />

similar events and also as a result <strong>of</strong> storm action (Lilley 1986:106). Lilley did not make<br />

any sampling-based inferences regarding archaeological distribution in the past, or use this<br />

type <strong>of</strong> information in inferring site formation processes, despite his Lapita-phase evidence<br />

(pottery from the KLK site on Tuam) being,<br />

“...associated with a very restricted quantity and range <strong>of</strong> other cultural<br />

material in a clean beach sand matrix which is devoid <strong>of</strong> structural features<br />

270


such as hearths and postholes (Lilley 1986:455).”<br />

He inferred, like Wickler for Buka, that the Lapita-phase pottery represented, “...probably<br />

only intermittent activity in Siassi generally and on Tuam in particular and second, that<br />

people did not actually live on the KLK site”, but did not regard the lack <strong>of</strong> occupation<br />

features as related to natural formation processes.<br />

<strong>The</strong> effect <strong>of</strong> storm surges and cyclone waves has been studied in some detail for<br />

Tongatapu archaeological sites (Spennemann 1987). Spennemann concluded that waves,<br />

including storm waves, could play an important part in site formation, site visibility, and<br />

site destruction, but gave examples also <strong>of</strong> beach scatters that were the transformed (lag<br />

deposit) remains <strong>of</strong> terrestrial archaeological sites. In his Tongan case study major<br />

erosional effects were limited to exposed sand cays lying <strong>of</strong> the northeast coast <strong>of</strong><br />

Tongatapu. He did not seek to develop any predictive use <strong>of</strong> the information, arguing<br />

instead for continual monitoring and rescue excavation <strong>of</strong> sites subject to coastal erosion.<br />

A general study on the subject <strong>of</strong> tropical cyclones and their effect on Pacific<br />

archaeological sites (Tarlton 1996) considered cyclone frequency for various island groups<br />

in the southwest Pacific and their impact on various island types. <strong>The</strong> study focused<br />

exclusively on terrestrial archaeological sites, while in the Roviana case the evidence<br />

suggests maritime site location in the past. Thus her discussion <strong>of</strong> increases in water level<br />

as having largely negative impacts on site preservation is largely inapplicable to stilt<br />

occupation over water in sheltered locations, where increases in water level during storms<br />

should theoretically protect archaeological deposits from swash processes.<br />

Tarlton also concluded that high islands with indented coastlines were likely to<br />

have sheltered areas relatively unaffected by cyclone waves (Tarlton 1996:108, 114), and<br />

that landforms resulting from cyclone activity or tsunami were readily identifiable in many<br />

instances (Tarlton 1996:118), both conclusions pertinent to discussion <strong>of</strong> Roviana<br />

intertidal formation processes. Tarlton considered that awareness <strong>of</strong> cyclone or other<br />

271


wave-generated landforms would enable an archaeologist to choose locations which were<br />

relatively undisturbed where stratified archaeological deposits might be discovered (Tarlton<br />

1996:123).<br />

Tarlton reviewed archaeological reports <strong>of</strong> cyclone or other wave damage to sites<br />

or landscapes. Among these the most sophisticated distributional inferences seem to be<br />

from Australia, for the Queensland coast (Rowland 1989), and the Torres Strait islands,<br />

where whole coastal landscapes have a lack <strong>of</strong> archaeological evidence thought to relate<br />

to wave damage <strong>of</strong> sites or wave erosion <strong>of</strong> coastlines (Vanderwal 1973:187). A similar<br />

explanation for the distribution <strong>of</strong> Lapita and other early sites in New Caledonia was cited<br />

as a personal communication from Jean-Christophe Galipaud in 1995 (Tarlton 1996:136-<br />

138).<br />

Tarlton provided an extensive section on Kirk’s numerical modeling <strong>of</strong> cyclone<br />

wave exposure for the Rarotongan coastline and suggested that cyclone wave impact on<br />

archaeological sites could be modeled in a similar way using GIS, perhaps in combination<br />

with remote-sensing identification <strong>of</strong> cyclone wave-generated landforms (Tarlton<br />

1996:189-196), but did not extend her discussion to the interpretation <strong>of</strong> regional<br />

archaeological distributions, seeing this approach as principally as an aid to site<br />

prospecting, for locating undisturbed archaeological sites. It is in this section though, that<br />

discussion <strong>of</strong> a predictive modeling approach for a specific coastline is closest to the type<br />

<strong>of</strong> analysis conducted in the present <strong>chapter</strong>. Her emphasis on storm surge models and<br />

wind models was appropriate for exposed coastlines and/or terrestrial sites, but these are<br />

inapplicable in the Roviana case, where collection sites comprise reworked intertidal and<br />

sub-tidal evidence rather than in-situ terrestrial sites.<br />

Roviana Intertidal Ceramics and Waves:<br />

272


Intertidal sites in the Roviana study were all thought to have had broadly comparable post-<br />

deposition natural formation processes, due to location in an almost landlocked lagoon<br />

setting bordered by either mainland high-island shorelines or upraised Plio-Pleistocene<br />

reefs. Discrimination between some <strong>of</strong> the models <strong>of</strong> formation processes constructed in<br />

Chapter 7 require some independent assessment <strong>of</strong> wave exposure variation between sites.<br />

A model <strong>of</strong> wave exposure is presented in this <strong>chapter</strong> which deals with seasonal weather<br />

patterns which can be expected to have prevailed in varying degrees in the past, and a<br />

model <strong>of</strong> wave formation based primarily on the variable “fetch”, which is the distance in<br />

a windward direction between a shore and the nearest land.<br />

Waves coming onto a beach increase in height and steepness until they break. As<br />

waves move into shallow water, circular orbits <strong>of</strong> the water particles become flattened, and<br />

some wave energy will be used in moving sediments to and fro on the sea bed. This is<br />

particularly so for shallow-sloped sea beds, such as the reef flats on which some <strong>of</strong> the<br />

Roviana ceramic sites are located. As a wave breaks, most <strong>of</strong> the energy is dissipated in<br />

the final mixing <strong>of</strong> water, sand and shingle (Open <strong>University</strong> Oceanography Team 1989:27-<br />

29). It is these two “swash zone” effects on sediments which are <strong>of</strong> interest to the shallow-<br />

water/ intertidal archaeologist. <strong>The</strong> generation <strong>of</strong> waves by seismic processes and by wind<br />

will be considered separately below.<br />

Tsunami:<br />

Exposure <strong>of</strong> shorelines within Roviana Lagoon to large waves generated by seismic events<br />

is comparatively low. <strong>The</strong> Blanche Channel, to which the Roviana reef passages open, is<br />

sheltered by Rendova/Tetepare islands, from tsunami originating in the Solomon Sea<br />

(Figure 6). Earthquake epicentres recorded between 1962 and 1967 show only a single<br />

shallow epicentre within the Blanche Channel, the vast majority <strong>of</strong> seismic events having<br />

their epicentres to the north, in the Bougainville area, or well south in the Southeast<br />

273


Solomons (Dunkley 1986:9, Mann et al. 1998). In the event that seismic activity did create<br />

waves in the Blanch channel, the effect within the Lagoon would be negligible due to the<br />

narrowness <strong>of</strong> the reef passages through the uplifted coral barrier islands. It is possible that<br />

refracted waves generated within the Blanche channel by seismic activity could have some<br />

impact on sites like Honiavasa and Miho, located at the inner end <strong>of</strong> a reef passage,<br />

adjacent to deeper water, but only a few small events <strong>of</strong> this type are known <strong>of</strong> , in contrast<br />

to their common occurrence on the Solomon Sea coasts <strong>of</strong> Rendova and elsewhere in the<br />

New Georgia group.<br />

<strong>The</strong> lagoon basin is interrupted with many linear and patch-like shoals, thought to<br />

indicate a submerged karst-like landscape (Stoddart 1969a:394), and generation <strong>of</strong><br />

substantial waves within the lagoon by seismic activity seems unlikely in view <strong>of</strong> the<br />

baffling effect <strong>of</strong> this dissected bathymetry.<br />

Wind-generated Waves:<br />

Waves in this context are defined as the process by which wind kinetic energy is<br />

transferred to and transported across water, as a progressive wave. For an idealised wave,<br />

wave height refers to the vertical distance between trough and peak, which is twice the<br />

wave amplitude, while wavelength is the distance between two successive waves.<br />

Steepness is wave height divided by wave length. Period is the time interval between two<br />

successive analogous points on the wave pattern passing a fixed location. <strong>The</strong> number <strong>of</strong><br />

peaks or troughs passing a fixed point per second is the frequency. Wave speed is related<br />

to wavelength (Open <strong>University</strong> Oceanography Team 1989:7, 10).<br />

On the open ocean, after a period <strong>of</strong> calm weather, when the wind starts to blow,<br />

small, steep waves form as the wind increases, continuing to grow after the wind has<br />

reached maximum speed until they reach a wavelength equal to one third <strong>of</strong> windspeed.<br />

Fetch is the unobstructed distance <strong>of</strong> sea in which the wave can build. Wave size depends<br />

274


on the fetch (Open <strong>University</strong> Oceanography Team 1989:11-12). Fetch imposes the prime<br />

limits on wave size, speed and energy in sheltered waters like Roviana lagoon and the<br />

Blanche Channel, and provides an easily measured proxy for wave exposure when<br />

combined with wind direction.<br />

“<strong>The</strong> factors that are important in increasing the amount <strong>of</strong> energy that<br />

waves obtain are (1) windspeed, (2), the time during which the wind blows<br />

in one direction, and (3) the fetch.....When both maximum fetch and<br />

duration are reached for a given wind velocity, the sea is said to be fully<br />

developed (Thurman 1981:219).”<br />

This is why windspeed and duration are less significant than fetch for the Roviana intertidal<br />

sites, as the limits imposed by short fetch are quickly reached, beyond which waves do not<br />

build. <strong>The</strong> following three anecdotes illustrate the primacy <strong>of</strong> fetch in determining wave<br />

exposure.<br />

During the 1998 cyclone (which tracked through the Rennell/Bellona area)<br />

northerly winds were estimated by the author to have exceeded fifty knots. In these wind<br />

conditions, children were sailing their canoes at respectable speed in the bay near the Miho<br />

site, powered by coconut fronds, across water barely more than rippled, due to the limits<br />

on wave energy imposed by short fetch across that bay. For a sea to become fully<br />

developed in similar windspeed to this would require a fetch <strong>of</strong> about 2500km and a<br />

duration <strong>of</strong> 65 hours (Thurman 1981:219).<br />

By way <strong>of</strong> contrast, moderate tradewind conditions in the Blanche Channel <strong>of</strong><br />

around 25 knots windspeed, <strong>of</strong> extended duration, where the fetch to the southeast is<br />

about 65km, allow a sea to become partially developed, with short and relatively steep<br />

waves hazardous to small craft a fetch <strong>of</strong> about 180km is needed for a sea to become fully<br />

developed for this windspeed).<br />

Wave refraction is well understood in oceanography, allowing oceanographers to<br />

determine regions that are likely to experience high waves due to this effect. Refraction<br />

275


can change wave height and direction, as can shallowing water (Open <strong>University</strong><br />

Oceanography Team 1989:25). Locations such as the inner margins <strong>of</strong> reef passages may<br />

be exposed to refraction waves even though the measured direct fetch is low.<br />

Seagrass, Mangroves and Reefs:<br />

Seagrasses, mangrove forests and reef shallows all serve to mitigate wave exposure, by<br />

effectively reducing fetch at low tide, acting as an energy absorbing barrier to wave<br />

formation and transmission.<br />

Roviana Lagoon nowadays has extensive shallow mixed seagrass meadows,<br />

characterized most visibly by the long strap-like Enhalus acoroides, in which Dugong<br />

dugon and Chelonia mydas graze, and which shelter a pr<strong>of</strong>usion <strong>of</strong> fish, mollusc and<br />

crustacean species. Besides being attractive locations for finding a meal (easily traversed<br />

by paddle canoe, and less easily if reliant on a rotating propellor), these meadows also have<br />

a sheltering effect on marine archaeological deposits, trapping organic and reefal detritus,<br />

and acting as sediment stabilizers (Stewart 1999:581). Seagrasses have the potential to<br />

create stratigraphy by sealing archeological deposits, and also are potentially agents <strong>of</strong><br />

bioturbation through the growth <strong>of</strong> root systems, as are many <strong>of</strong> the creatures that live<br />

within the shelter <strong>of</strong> seagrasses, some <strong>of</strong> which are burrowers and some <strong>of</strong> which pluck<br />

grasses from sediments when grazing.<br />

Seagrass shallows are found principally in the eastern two-thirds <strong>of</strong> the Lagoon,<br />

east <strong>of</strong> Honiavasa, where the lattice <strong>of</strong> shallow reefs nears the surface <strong>of</strong> the water, as a<br />

result <strong>of</strong> a slight tectonic tilting (Mann et al. 1998). West <strong>of</strong> Miho site, the coral heads tend<br />

to be deeper below the surface, as far west as Nusa Roviana, although there is a lattice <strong>of</strong><br />

shallow reefs along the northern (mainland) coastline <strong>of</strong> the lagoon in this region.<br />

Seagrasses are not common west <strong>of</strong> Miho, other than sparse occurrences along the shallow<br />

mainland coast and among the forest <strong>of</strong> islets between Miho and Zangana.<br />

276


Forty-three percent <strong>of</strong> the worlds mangrove species are represented in Solomon<br />

Islands, with 26 species in 13 families (Oreihaka 1997). Mangroves protect many <strong>of</strong> the<br />

Lagoon’s shorelines from wind and waves, having a similar stabilizing role to seagrasses,<br />

and harbouring crustaceans, molluscs and many different fish species, some <strong>of</strong> which are<br />

agents <strong>of</strong> bioturbation. Mangroves may create sedimentary strata post-deposition. At the<br />

Hoghoi site, for example, testpits revealed an orange-brown organic layer beneath the top<br />

10cm <strong>of</strong> sediment, composed almost entirely <strong>of</strong> living (mangrove?) rootlets, with sherds<br />

found both above and below this layer. This could end up looking like bedding planes in<br />

a deposit <strong>of</strong> pottery fragments after the death <strong>of</strong> the mangrove system, which would leave<br />

a thin dark organic band in an excavation section pr<strong>of</strong>ile. Mangroves are found principally<br />

along the mainland coast <strong>of</strong> the lagoon, where s<strong>of</strong>t sediments predominate, and are present<br />

in smaller areas along parts <strong>of</strong> the raised-coral barrier system, where they are found<br />

associated with s<strong>of</strong>ter sediments.<br />

Solomon Islands Tides:<br />

<strong>The</strong> reason tides are significant is that land and sea breezes follow a diurnal cycle, as do<br />

tides in the region for much <strong>of</strong> the year, thus creating a situation in which wind directional<br />

changes or changes in wind strength are in phase with tidal changes for extended periods,<br />

as will be detailed below, leading to differential wave exposure <strong>of</strong> coastlines in some<br />

seasons. Numerous shallow reefs within the lagoon shelter many <strong>of</strong> the shorelines at low<br />

tide (when intertidal scatters are most at risk <strong>of</strong> wave damage) but for those shorelines not<br />

sheltered by reefs within the Lagoon, sea-breeze wave exposure at low tide is thus likely<br />

to have had a more severe effect on the pottery scatters, which are more exposed to<br />

swash-zone energy at low tide.<br />

Lunar-induced tides have a fundamental periodicity <strong>of</strong> 12 hours and 25 minutes,<br />

277


and operate over a cycle <strong>of</strong> 27.3 days governed by the rotation <strong>of</strong> the moon-earth system.<br />

Usually, in most parts <strong>of</strong> the world, for this reason, the time <strong>of</strong> the high tide shifts slightly<br />

each day (the diurnal calendar being based on a 24 hour solar periodicity). Lunar tidal<br />

forces are also influenced by the moon’s declination either side <strong>of</strong> the equatorial plane,<br />

which varies over a cycle <strong>of</strong> 27.2 days, and controls diurnal variation. <strong>The</strong> Moon’s<br />

elliptical orbit also causes variation in tide-producing forces over a 27.5 day cycle<br />

(Macmillan 1966:48-49). Solar tidal forces are <strong>of</strong> smaller magnitude than lunar tidal forces,<br />

with a semi-diurnal period <strong>of</strong> twelve hours.<br />

<strong>The</strong> declination <strong>of</strong> the sun varies over a yearly cycle, and the elliptical orbit <strong>of</strong> the<br />

earth around the sun has a slight effect on a yearly cycle also. Diurnal inequalities (the<br />

tendency for the two tides <strong>of</strong> a 25 hour period to be <strong>of</strong> unequal height) are affected by<br />

declination <strong>of</strong> the sun and moon, and will be greatest at times <strong>of</strong> maximum declination, and<br />

least at times <strong>of</strong> minimal declination (Macmillan 1966:44). This diurnal inequality is<br />

significant for the Roviana case, where for much <strong>of</strong> the year a solar tidal cycle<br />

predominates, as will be explained below. When solar and lunar effects are in phase, either<br />

in conjunction or opposition, large tidal ranges occur, and when they are out <strong>of</strong> phase the<br />

range is small. <strong>The</strong>se are referred to as spring tides and neap tides respectively, changes<br />

which follow a 29.5 day cycle (Macmillan 1966:50).<br />

<strong>The</strong> Pacific ocean appears to favour the occurrence <strong>of</strong> small-range diurnal tides,<br />

and furthermore, the solar contribution sometimes exceeds that <strong>of</strong> the lunar in the Pacific<br />

area,<br />

“...probably due to the resonance <strong>of</strong> the solar disturbing period with the<br />

natural period <strong>of</strong> the body <strong>of</strong> water affected. In consequence, under these<br />

circumstances the diurnal tide due to solar influence dominates that due to<br />

the moon (Macmillan 1966:60).”<br />

Solomon Islands tides seem to be <strong>of</strong> the mixed type, varying seasonally between diurnal<br />

and semi-diurnal with the changing influence <strong>of</strong> solar forces. Tide predictions for Honiara<br />

in year 2000 and 2001 (Sea levels for Honiara were obtained from the National Tidal<br />

278


Facility, <strong>The</strong> Flinders <strong>University</strong> <strong>of</strong> South Australia) show a mixed tide regime where a<br />

diurnal tidal cycle (1 high tide every 24 hours) is dominant for most <strong>of</strong> the year, with high<br />

water occurring near noon in January, February and early march, tending to have a<br />

morning high water in March, and shifting to a semi-diurnal regime in April (two tides in<br />

24 hours), with only small tidal range for this cycle. May sees a return to a diurnal cycle,<br />

but with low tide during the day and high tide in the early hours <strong>of</strong> the morning. June is<br />

dominated by a diurnal cycle with low tide at mid-day and high water at night, and this<br />

pattern continues through July and August, being modified towards a semi-diurnal cycle<br />

in September and October. November sees a return to a stronger diurnal cycle, but with<br />

high water during the day, with the same pattern in December. Although tidal records for<br />

the New Georgia group were not accessed, this pattern is consistent with the author’s own<br />

observation and information from local informants during 1996-1998 at Roviana Lagoon.<br />

This annual pattern shows a strong influence <strong>of</strong> solar declination, dividing the year<br />

into four tidal seasons, with the sun creating a semi-diurnal effect at the equinoxes and a<br />

diurnal regime at the solstices. Within this annual pattern is a monthly cycle created largely<br />

by the changing declination <strong>of</strong> the moon, and the rotation <strong>of</strong> the moon-earth system. <strong>The</strong>re<br />

is a daytime high tide from November to March (October to January according to Aswani<br />

1997:235), and a night time high tide from May through September. <strong>The</strong> timing <strong>of</strong> daily<br />

high-water is following a solar cycle rather than a lunar one (as outlined by Macmillan<br />

1966:60).<br />

<strong>The</strong> daytime solar tide is superior around the summer solstice, and the night-time<br />

tide becomes superior around the time <strong>of</strong> winter solstice. Why this should be in the daytime<br />

through the months <strong>of</strong> northern solar declination, and at night through the months <strong>of</strong><br />

southern declination is explained by a the mechanism <strong>of</strong> diurnal inequality (Macmillan<br />

1966:42), whereby the declination <strong>of</strong> the relevant tide-producing body (in this case the sun)<br />

produces, instead <strong>of</strong> the theoretical equilibrium tide, superior and inferior semi-diurnal<br />

tides (in the Roviana case the inferior tide is completely suppressed when the declination<br />

<strong>of</strong> the moon is low, but is present as a slight tidal oscillation when declination <strong>of</strong> the moon<br />

is high, on a monthly cycle).<br />

<strong>The</strong> broad characteristics <strong>of</strong> this tidal system as explicated above seem likely to be<br />

a pattern <strong>of</strong> antiquity as well as the present, as there seems to be few reasons for this<br />

to have varied over the timescale <strong>of</strong> the past few thousand years. Only at the geological<br />

timescale, where the Pacific Ocean basin undergoes major changes in plate tectonics, can<br />

this resonance with solar tidal forces be expected to change, as a result <strong>of</strong> the changing<br />

279


size and depth <strong>of</strong> the basin, within the current (and longstanding) harmonic theory <strong>of</strong> tides.<br />

Winds:<br />

<strong>The</strong> nearest meteorological station to the Roviana Lagoon is situated in Munda, away from<br />

the coast, surrounded by trees and buildings, and while this location is no doubt convenient<br />

for staff who take the readings, recorded windspeeds and directions cannot be expected<br />

to bear much relationship to winds affecting the Lagoon. This section is therefore written<br />

using a mix <strong>of</strong> data pertaining to the general region acquired from the Pacific Pilot<br />

(Hydrographic-Office 1971: diagrams 8-11), data from personal observation, and data<br />

from Aswani’s thesis (Aswani 1997:238-239). (Author’s note: the airport building was<br />

shifted circa 2001, allowing slightly more exposed location for weather equipment.)<br />

<strong>The</strong> southeast trades (Gevasa) prevail from April/May/June through to about<br />

October, and their strength and duration varies from year to year. Direction is recorded as<br />

blowing from 125 degrees true (Hydrographic-Office 1971:diagram 10). Personal<br />

observation over the course <strong>of</strong> three years suggests the local direction <strong>of</strong> this wind is more<br />

like 135 degrees, as at the Honiavasa reef passage it blows consistently directly from<br />

Tombatuni Island in the Blanche channel. This wind tends to strengthen during the day,<br />

probably due to the local effect <strong>of</strong> the sea breeze, at a time <strong>of</strong> year when tides are low<br />

during the day, so that any pottery scatter at such a depth as to be exposed to the<br />

Southeast, or to refraction waves from the Blanche channel curving around the inner points<br />

<strong>of</strong> the reef passages, can be expected to have undergone a relatively harsh taphonomic<br />

process. Low tides during the day during the season <strong>of</strong> the SE trade mean that any<br />

intertidal pottery sites with a large fetch to the SE (e.g. Gharanga, or the Mbolave lithic<br />

scatter/ceramic findspot) can be expected to undergo a harsh swash-zone taphonomic<br />

regime, unless sheltered by reefs or s<strong>of</strong>t sediments and vegetation.<br />

<strong>The</strong> Northeast trade blows strongly for a period <strong>of</strong> a month or so in late summer,<br />

but locally from the Northwest, and is known by local expatriates as the Northwest<br />

Monsoon (this may be the “Peza” referred to by Aswani, but the period and direction<br />

differ from Aswani’s suggestion that this wind is a westerly. <strong>The</strong> Roviana dictionary lists<br />

Peza as the northwest monsoon (Waterhouse 1926:91). <strong>The</strong> recorded direction <strong>of</strong> origin<br />

<strong>of</strong> the NE trade is about 70 degrees true (Hydrographic-Office 1971: diagram 9). Personal<br />

observation suggests that direction is locally influenced by the New Georgia landmass<br />

280


with its volcanic peaks, tending to blow more from the north to NNW, and tending to<br />

strengthen in the evening and at night with the addition <strong>of</strong> a slight land breeze. Tides at this<br />

time <strong>of</strong> year are predominantly semi-diurnal (two highs and two lows every 24 hours), so<br />

are not in phase with this strengthening <strong>of</strong> the wind in the evening. Sites exposed to<br />

significant Northwesterly fetch in this direction at low tide can be expected to show some<br />

effect <strong>of</strong> this harsher taphonomic regime.<br />

Transitional periods between these seasonal prevailing winds are dominated by a<br />

diurnal cycle <strong>of</strong> light land (night) and sea (day) breezes (Sea breezes blow in from the<br />

Blanche Channel toward the centre <strong>of</strong> New Georgia during the day, and land breezes blow<br />

<strong>of</strong>f the coast towards the Blanche Channel at night). <strong>The</strong>se breezes are most noticeable at<br />

midsummer, in the absence <strong>of</strong> other winds, but the forces generating these winds, the<br />

differential heating and cooling <strong>of</strong> land and sea, act in combination with other winds in the<br />

cooler months also. When tides are high during the day, sites with seaward fetch are<br />

unlikely to be affected as the pottery is well submerged, but sites with landward fetch at<br />

low tide at night (Hoghoi for example) might be expected to be exposed to a build-up <strong>of</strong><br />

small waves through the course <strong>of</strong> the night.<br />

Tropical depressions, tropical storms, and cyclones tend to originate between<br />

Latitude 10 degrees south and Latitude 15 degrees south (Hydrographic-Office 1971:36<br />

and diagram 12). <strong>The</strong>re are an average <strong>of</strong> about two cyclones per year in the region, and<br />

these generally pass to the south <strong>of</strong> the Solomons, and can be expected in most instances<br />

to create strong winds veering (progressing around the compass clockwise) through<br />

northerly as a result although other directions are not ruled out, as some rare cyclones<br />

originate in Solomon Islands. Holocene climate changes may have affected cyclone tracks<br />

and other climatic zonation in the past (Tarlton 1996:18-28), but no information is<br />

available at present on the nature <strong>of</strong> such changes in the period since deposition <strong>of</strong> the<br />

pottery. It seems likely that the presence <strong>of</strong> the high landmasses <strong>of</strong> Choiseul and Ysabel<br />

to the east are moderating influences restricting the formation <strong>of</strong> cyclones in the region,<br />

and that this effect is <strong>of</strong> great geological age.<br />

<strong>The</strong> season for tropical storms is December through April, with most occurring in<br />

January, February and March. As their average track speed is about eight knots, they can<br />

affect an area for many days and nights (Hydrographic-Office 1971:36), so tidal periodicity<br />

is irrelevant to wave exposure during tropical cyclones.<br />

Cyclone storm surge may (but does not necessarily) raise sea levels temporarily,<br />

reducing the effect <strong>of</strong> waves raised during a cyclone on the intertidal pottery scatters,<br />

281


elative to effects <strong>of</strong> lesser winds prevailing at times <strong>of</strong> low tide. Positive tidal surges <strong>of</strong> up<br />

to 4m occur during tropical cyclones (Open <strong>University</strong> Oceanography Team 1989:61), and<br />

in some cases this property <strong>of</strong> tropical storms and cyclones would be expected to reduce<br />

the effect <strong>of</strong> storm generated waves in the lagoon on submerged pottery scatters. Storm<br />

surge is caused primarily by winds driving water ashore and this will depend at a larger<br />

scale on whether wind is onshore or <strong>of</strong>fshore, but within the “radius <strong>of</strong> maximum wind”,<br />

several kilometres from the storm centre, surge response to barometric pressure anomaly<br />

is universal for cyclones (Anthes 1982:5-7, 161), although this effect is slight. Where<br />

cyclones pass to the south <strong>of</strong> the Solomon Islands, as is usually the case, and winds can be<br />

expected to veer through northerly, storm surge would be expected to be slight or absent,<br />

and was observed to be slight during such an event in 1998, but if a cyclone were to take<br />

a more northerly route, storm surge from Easterlies, Southerlies or Westerlies could be<br />

expected in the embayment comprising the Blanche Channel. In terms <strong>of</strong> cyclone events,<br />

the circumstance most likely to affect intertidal pottery scatters would be the more<br />

common northerly-tending cyclone winds with attendant storm surge only slight or absent,<br />

as pottery might be exposed to wave action in the lagoon at low tide in these<br />

circumstances. Sites with significant fetch in a northerly direction (NE through to NW)<br />

could therefore be expected to be more severely impacted by cyclones.<br />

Measuring Fetch:<br />

Fetch for cyclones was measured as the maximum at low tide in a northerly arc between<br />

NW and NE. As the direction <strong>of</strong> the NE trade is uncertain on present information, fetch<br />

for this wind is measured as a maximum between 320 and 350 degrees (consistent with<br />

personal observation) at low tide; and also at 70 degrees at low tide, consistent with the<br />

direction <strong>of</strong> the NE trade shown in the Pacific Pilot. Fetch for the SE trade is measured at<br />

low tide bearing 140 degrees, consistent with personal observation. Fetch was not<br />

measured for the summer sea breeze because it coincides with the diurnal summer-solstice<br />

daytime solar tide, and is unlikely to affect the site areas studied. Fetch in the direction <strong>of</strong><br />

the land breeze was measured as the maximum at low tide 25 degrees either side <strong>of</strong> North.<br />

Fetch was measured using air photographs at a scale <strong>of</strong> 1:33 000.<br />

282


Results:<br />

Fetch under various combinations <strong>of</strong> wind and tide are given in Table 21.<br />

Table 21: Fetch measurements for collection sites as an indicator <strong>of</strong> wave exposure.<br />

Site Cyclone<br />

(NW-NE)<br />

Paniavile 2km<br />

(moderated<br />

by seagrass<br />

and s<strong>of</strong>t<br />

sediments)<br />

Hoghoi 6km<br />

(seagrass)<br />

“NE” trade<br />

(Peza?)<br />

283<br />

SE trade<br />

(Gevasa)<br />

land breeze<br />

(low night)<br />

Maximum<br />

fetch (max<br />

exposure)<br />

0 0


pottery discard is less likely to have survived.<br />

Hoghoi had a wave exposure measurement <strong>of</strong> up to six km to the NNE, with wave<br />

effects moderated due to protection at low tide by an extensive shallow seagrass flat with<br />

numerous small coral heads nearing the surface. If sea levels were higher in the past this<br />

protection would have been absent or reduced, but the ceramics would have been in deeper<br />

water, mitigating swash effects to some extent.<br />

Gharanga could have up to six km exposure in SE tradewind conditions allowing<br />

swash to affect the site at intermediate stages <strong>of</strong> the tide (this effect may have been<br />

extended in the past if sea levels were slightly higher, with the materials in shallow water<br />

rather than exposed at low tide). Gharanga and Kopo are different to the other sites in that<br />

they are situated at the mouth <strong>of</strong> mainland streams. Sherds did not have the cushioning<br />

protection <strong>of</strong> marine growth to the same extent due to reduced growth <strong>of</strong> sponges in the<br />

low-salinity location, and could also be subject to taphonomic effects arising from stream<br />

flooding.<br />

Miho and Honiavasa both yielded maximal wave exposure <strong>of</strong> 6km but Honiavasa<br />

was more exposed to refraction waves in SE tradewind conditions (Miho is better<br />

protected from such waves than Honiavasa due to more extensive shoals adjacent to the<br />

passage and situation in a small embayment, while Honiavasa, by contrast, is located quite<br />

close to the deeper water <strong>of</strong> the passage, and the author noted refracted waves while<br />

collecting there on a rising tide in tradewind conditions. While the seaward portions <strong>of</strong> the<br />

site were in deep enough water to be protected in these conditions, shoaling towards the<br />

centre <strong>of</strong> the site left the latter area exposed to swash, while the eastern end <strong>of</strong> the site was<br />

protected by the central shoal. Sediments across the site support the idea <strong>of</strong> a central area<br />

affected by swash from refraction waves, where ceramics are exposed as large fragments,<br />

and a more sheltered easterly zone where ceramic visibility drops away as sandier<br />

sediments build up. <strong>The</strong> picture could look quite different after a strong northerly though.<br />

284


Nusa Roviana has up to 3km <strong>of</strong> fetch in a NNE direction, and can be expected to<br />

show some signs <strong>of</strong> damage or size sorting in view <strong>of</strong> this relatively exposed location (in<br />

modern times a wharf/breakwater has been constructed here to shelter the landing in<br />

northerly conditions). Sherd surfaces were generally quite damaged in this site with temper<br />

grains slightly pedestalled in many cases (the single dentate sherd was an exception,<br />

retaining a smooth finish).<br />

Unanalysed assemblages from Ririgomana, Punala, Humbi Kongu and Rango were<br />

all in poor condition with sherds showing signs <strong>of</strong> water rolling in many cases. More<br />

recently, large sherds in good condition have been reported from the deeper margins <strong>of</strong><br />

Rango (Rigeo, Pers. comm., 2002). Unlike Paniavile, Ririgomana is free <strong>of</strong> seagrass, and<br />

although the water is shallow generally in the area between Ririgomana and the mainland,<br />

I can envision a chop <strong>of</strong> small waves building up along this shore in a northerly. Punala and<br />

Humbi Kongu were similar in this respect.<br />

A number <strong>of</strong> findspots and lithic scatters on the rocky shore <strong>of</strong> Honiavasa Island<br />

to the east <strong>of</strong> Hoghoi had a lesser degree <strong>of</strong> seagrass shelter than Hoghoi, as attested also<br />

by rockier coral shorelines. <strong>The</strong> low-density scatter <strong>of</strong> sherds, especially lithic manuports,<br />

along this shoreline does not preclude relatively intensive stilt-house settlement in the past,<br />

with most ceramics removed by swash-zone rolling post-deposition. Similarly, the ceramic<br />

findspot at Kazu was on a stony shoreline, and may be a chance survival <strong>of</strong> significant SE<br />

trade exposure. Lithic manuport scatter at Mbolave, where a single sherd was found, could<br />

also be a case <strong>of</strong> a ceramic/lithic scatter transforming into a lithic scatter through time.<br />

<strong>The</strong> fetch data predicts that Honiavasa, Miho, Gharanga and Hoghoi should show<br />

the greatest level <strong>of</strong> sherd damage and size-sorting, closely followed by Nusa Roviana.<br />

When the sheltering effect <strong>of</strong> seagrass shallows is taken into account at Hoghoi, this<br />

285


seems to partly match the size-data from sites (Table 22) in that Honiavasa and Miho had<br />

larger average sherd sizes than other sites (the 25 sherds from Kopo were largest but this<br />

may be a sample size and/or collection intensity effect).<br />

Table 22: Total sherd count and average sherd area by collection site,<br />

resulting from combined effects <strong>of</strong> collection intensity and wave exposure.<br />

Site Sherd Count Average Sherd<br />

Area(cm 2 )<br />

Paniavile 644 14.65<br />

Hoghoi 861 10.94<br />

Miho 382 22.08<br />

Honiavasa 442 28.25<br />

Gharanga 277 17.96<br />

Nusa Roviana 115 15.83<br />

Kopo 25 33.24<br />

Zangana 860 13.59<br />

Chapter Summary and Conclusions:<br />

It was difficult to construct an unequivocal model <strong>of</strong> wave exposure given conflicting<br />

information about the direction <strong>of</strong> prevailing winds, and uncertainty as to the relative<br />

importance <strong>of</strong> the various winds. Personal observation and local anecdotal information<br />

conflict with the information in the 1971 Pacific Pilot, and the latter is probably only<br />

intended as a very general guide for navigational use. Aswani’s recording <strong>of</strong> Peza as a<br />

westerly may be a local effect due to topography at Baraulu Village (Canaan), where he<br />

was mostly based, and a northwesterly direction for this strong wind is preferred here.<br />

Wave exposure for some sites (e.g. Gharanga and Hoghoi) is much higher using<br />

286


Figure 75: Location <strong>of</strong> Honiavasa and Miho collection sites<br />

at the inner ends <strong>of</strong> a long passage through uplifted barrier<br />

reef. Miho is more sheltered from refraction waves than<br />

Honiavasa as it is situated in a small bay oriented to the<br />

NW, while the western edge <strong>of</strong> Honiavasa drops away into<br />

Honiavasa passage.<br />

personally observed wind directions, and any future modeling <strong>of</strong> this sort would benefit<br />

from more formal recording <strong>of</strong> wave exposure at various locations in the various prevailing<br />

conditions.<br />

This analysis provides a caution against a direct reading <strong>of</strong> aceramic scatters as<br />

having a chronological or behavioural significance, suggesting that some <strong>of</strong> these at least<br />

may be the result <strong>of</strong> a harsh taphonomic regime, which has biased the observed assemblage<br />

composition towards relatively robust coarse-grained lithic manuports.<br />

Kopo and Paniavile seemed to be the more sheltered locations, and this seems<br />

consistent with the information initially recorded regarding Paniavile, which was a rich site<br />

prior to a series <strong>of</strong> collecting and fossicking events (Reeve 1989). <strong>The</strong> sample <strong>of</strong> sherds<br />

from Kopo is too small to yield useful taphonomic analysis results, but collection intensity<br />

was fairly high (Sheppard, pers. comm. 2002) suggesting either a harsher taphonomic<br />

regime than predicted by the wave exposure model or a low intensity-duration <strong>of</strong><br />

287


occupation in the past. It seems likely that a number <strong>of</strong> such sites may be well preserved<br />

along the mainland shoreline under s<strong>of</strong>t sediments, in locations sheltered from all the major<br />

wind/fetch combinations (the sort <strong>of</strong> locations which can be expected to be characterized<br />

by a buildup <strong>of</strong> s<strong>of</strong>t sediments). <strong>The</strong> reported ceramic find-spot at Pikoro may be one such<br />

location, although limited test-pitting to the water table revealed a sparse distribution <strong>of</strong><br />

lithic manuports and some small sherd fragments from the adjacent stream banks.<br />

None <strong>of</strong> the ceramic sites had more than 6km maximum fetch, suggesting this may<br />

be an upper limit for site preservation in the water depths in which the sites were found<br />

(intertidal-subtidal). This analysis thus provides a starting point for designing a model <strong>of</strong><br />

site preservation in relation to wave exposure for this type <strong>of</strong> site. <strong>The</strong> model could be<br />

extended by adding data from other areas where similar sites are found at different depths.<br />

This sort <strong>of</strong> modeling process is fundamental to understanding the distribution <strong>of</strong> Lapita<br />

across Near Oceania. <strong>The</strong> Reef-Santa-Cruz terrestrial Lapita sites at the near margin <strong>of</strong><br />

Remote Oceania and terrestrial sites in the Talasea region may mark a shift by some early<br />

Lapita-era people onto land.<br />

288


CHAPTER 7:<br />

SITE/ASSEMBLAGE FORMATION PROCESSES<br />

Introduction:<br />

(SHERD TAPHONOMY).<br />

Determination <strong>of</strong> formation processes (or at least thorough consideration <strong>of</strong> the<br />

possibilities) is widely regarded as an essential stage in making archaeological inferences,<br />

although,<br />

“Many archaeologists simply ignore the subject, perhaps because they are<br />

unconvinced <strong>of</strong> its importance or wary <strong>of</strong> its implications for the validity<br />

<strong>of</strong> their interpretations <strong>of</strong> the past (Shott 1998).”<br />

Interest in archaeological formation processes burgeoned during the 1970s (Schiffer<br />

1995c), however, for archaeological sites located in the sea, development <strong>of</strong> formation<br />

theory and method has lagged by comparison (Stewart 1999). A key question facing the<br />

analyst is whether comparisons between samples, for example, seriation, are invariant<br />

under the transformations brought about by site formation processes (Orton 2000:66). For<br />

the Roviana intertidal sites, another fundamental behavioural question is whether<br />

deposition was in the sea in the past, or whether submergence/coastal erosion has made<br />

intertidal sites out <strong>of</strong> terrestrial sites. Although marine transgression is primarily a<br />

destructive process (Waters 1992:275), it is clear that coastal occupation sites can survive<br />

submergence relatively unscathed in ideal circumstances, these being where sites are<br />

sheltered from wave energy (Flemming 1983) or where submergence (and sometimes<br />

burial in sediments) is rapid (Stewart 1999:572, Waters 1992:277), as in the case <strong>of</strong> Port<br />

Royal (Hamilton & Woodward 1984), and sufficiently deep to evade wave action (Waters<br />

1992:250-251). Comparison <strong>of</strong> Roviana density <strong>of</strong> deposits with the quantities <strong>of</strong> sherdage<br />

289


expected from estimates <strong>of</strong> the breakage population (Felgate 2002, Felgate & Bickler n.d),<br />

as discussed in Chapter 5, suggests that Roviana sherd samples are far from pristine, and<br />

that substantial sherd attrition has occurred during formation (see Chapter 6). <strong>The</strong><br />

possibility exists that the sites originated as terrestrial deposits and have been damaged in<br />

the course <strong>of</strong> becoming intertidal deposits.<br />

I will argue below against submergence or coastal erosion <strong>of</strong> terrestrial sites as an<br />

explanation for the observed archaeological distribution. My arguments in doing so closely<br />

parallel those <strong>of</strong> Reeves, Specht and Wickler for similar sites elsewhere in Near Oceania<br />

(Reeve 1989, Specht 1991, Wickler 2001), and follows closely the argument to this effect<br />

presented previously for Roviana underwater sites (Felgate 2002). Recent paleoshoreline<br />

research is cited as additional support <strong>of</strong> this argument (Mann et al. 1998).<br />

In this <strong>chapter</strong> taphonomic approaches to the identification <strong>of</strong> formation processes<br />

are taken, seeking to determine, from the characteristics <strong>of</strong> artefact assemblages and<br />

individual artefacts, the processes that have acted on these in forming the observed sites.<br />

A series <strong>of</strong> models <strong>of</strong> cultural formation processes are presented below, followed by a<br />

series <strong>of</strong> post-depositional formation process models. A range <strong>of</strong> sherd-based taphonomic<br />

analyses contribute to discriminating between these various formation models. In Chapter<br />

11 a spatial approach to the identification <strong>of</strong> postdepositional process is presented, while<br />

in Chapter 6 wave exposure was modelled for sites, and compared with other taphonomic<br />

information in this <strong>chapter</strong>.<br />

Beyond the identification <strong>of</strong> formation processes, a second major aim in seeking<br />

to understand taphonomic effects on assemblage composition is to assess the likely nature<br />

<strong>of</strong> taphonomic bias in assemblage composition. Assemblage stylistic composition is an<br />

important data source for chronological inference, through seriation analysis (see Chapter<br />

12). While it does not seem possible to fully understand the exact nature <strong>of</strong> assemblage<br />

biases arising through taphonomic processes, a number <strong>of</strong> observations can be made that<br />

inform in this regard, and which aid in the analysis <strong>of</strong> decorative variability.<br />

290


Review: Middle Range <strong>The</strong>ory for the Identification <strong>of</strong> Formation Processes:<br />

A general review <strong>of</strong> suggested methods by which formation processes might be identified<br />

is given here. <strong>The</strong> classic review is that <strong>of</strong> Schiffer (1995c), whose categories <strong>of</strong> evidence<br />

are used here.<br />

Size effects, density, shape, orientation, structure and context <strong>of</strong> the deposit:<br />

<strong>The</strong>re is a substantial literature on the size-sorting effects <strong>of</strong> cultural transformation<br />

processes such as cleanup activities in the past. <strong>The</strong> McKellar principle, for example<br />

states,<br />

“ that small artefacts, especially microartefacts on occupation surfaces<br />

<strong>of</strong>ten indicate primary refuse, whereas in activity areas not habitually<br />

cleaned, such as some lithic workshops, abandoned structures... and vacant<br />

lots..., larger items can accumulate as primary refuse (Schiffer<br />

1995c:175).”<br />

Perhaps more important than such behavioural correlates in the present context are non-<br />

cultural size-sorting effects as governed by the laws <strong>of</strong> hydrology or hydraulic transport,<br />

viewing archaeological items as sedimentary particles (e.g. Shackley 1978). <strong>The</strong>re are<br />

many studies dealing with stream and river environments from this perspective (Waters<br />

1992:321), and also hydrological experimental sedimentological approaches such as flume<br />

experiments (e.g. Day 1980). For coastal sediments, wave processes assume a greater<br />

importance than currents, as waves entrain sediments in a zone extending from the wave<br />

base to the landward margin <strong>of</strong> the swash zone, sediments which are then transported by<br />

beach drift, longshore current, and rip currents (Waters 1992:249-251). <strong>The</strong>se<br />

erosive/transportation processes fluctuate in magnitude during changes in weather, and are<br />

displaced landward and seaward with tidal fluctuations. When such coastal high-energy<br />

processes come into contact with archaeological sediments, the fine fraction <strong>of</strong> the<br />

artifact/sediment matrix becomes suspended in the water column, while a lag deposit <strong>of</strong><br />

291


larger/heavier items may be abraded by suspended sand or reworked by rolling (Waters<br />

1992:270-271).<br />

<strong>The</strong>se processes can create a swash-zone lag deposit by eroding and transporting<br />

terrestrial sediments and artefacts, or can rework materials initially deposited in the water<br />

as primary refuse. Both terrestrial and in-water discard have similar consequences for the<br />

size, shape, density and orientation <strong>of</strong> artefacts in some respects. Artefacts in such settings<br />

should occur in a matrix <strong>of</strong> similarly large or dense sedimentary particles. Some shapes will<br />

transport more easily by benthic rolling than others, while flatter shapes might be more<br />

easily suspended in the water column by turbulence than blockier shapes. <strong>The</strong> same across-<br />

shore size/shape/density-sorting evident in the sedimentary particles should show up in the<br />

size patterning <strong>of</strong> artefacts. Armouring <strong>of</strong> graded sediments may also occur, where a<br />

surface layer <strong>of</strong> larger, less easily transported items protects less-sorted underlying<br />

sediments and artefacts.<br />

To distinguish between erosion <strong>of</strong> a terrestrial deposit into a maritime lag deposit<br />

and primary deposition in the sea, one must look not only at the lag deposit itself, but<br />

landward, in search <strong>of</strong> traces <strong>of</strong> a source <strong>of</strong> the artefacts (remembering also that storm<br />

wash can transport materials from sea to land). This is the line <strong>of</strong> reasoning used by others<br />

in suggesting reef flat lithic/ceramic scatters represent deposition in the sea in the past<br />

(Reeve 1989, Specht 1991, Wickler 2001: 241). Absence <strong>of</strong> adjacent terrestrial erosion<br />

features that might supply a source <strong>of</strong> artefacts would suggest that the materials were<br />

deposited in the sea directly by some means.<br />

In deep water, deposition could occur as a result <strong>of</strong> a shipwreck or capsize, or<br />

funerary/religious practices, or discard from anchored vessels (Stewart 1999). In the<br />

intertidal zone or on the reef flat, one would look shoreward in search <strong>of</strong> an attractant that<br />

would cause a concentration <strong>of</strong> activity in such shallow water, such as a local resource (for<br />

example water or gardening land) that might cause intensive canoe traffic and an<br />

accumulation <strong>of</strong> pottery breakage events and discard. If there is a widespread pattern<br />

292


comprising many instances <strong>of</strong> reef-flat or intertidal site locations, without attendant<br />

deposits on land, and without obvious localized attractant resources, this is more likely a<br />

pattern resulting from settlement. Artifact accumulation quantity/density and assemblage<br />

composition/diversity can allow some further idea <strong>of</strong> site function and occupation intensity<br />

in such circumstances.<br />

Artefact size/density/shape should give clues here, even if the shore deposits have<br />

been reworked by storm wash. In the case <strong>of</strong> primary deposition on land, the terrestrial<br />

artefact deposit should show less size/density sorting than the adjacent marine lag deposit,<br />

as it cannot have been reworked to the same extent as the materials in the sea. In the<br />

circumstance where the artefacts on land are rounded and small, while the artefacts in the<br />

sea are <strong>of</strong> a greater variety <strong>of</strong> sizes and in good condition, one must conclude either that<br />

some harsh taphonomic process such as gardening has reduced sherds on land, or that the<br />

small sherds on land are a wave-transported size fraction that originated in the sea. Even<br />

in the circumstance where sherds on land are not particularly well sorted by size, there is<br />

the possibility that a more random size-selection has been cast up from the sea by large<br />

waves, and that only some lucky sherds still in the sea have been sheltered from the force<br />

<strong>of</strong> the water by microtopography. In such circumstances other geomorphological<br />

characteristics <strong>of</strong> storm wash deposits aught to be identifiable (for example Gosden &<br />

Webb 1994).<br />

Processes affecting artefact orientation and dip are not particularly well understood<br />

as yet. Dip may be particularly useful for identifying primary deposition <strong>of</strong> artefacts<br />

tramped onto a flat surface, and orientation has been used to infer aeolian and fluvial<br />

processes (Schiffer 1995c:178). Marine currents and some propwash effects from treasure-<br />

hunting techniques may leave patterned artefact orientation (Stewart 1999). Strong<br />

currents may occur at times in reef-flat locations near reef passages.<br />

Use Life, Damage and Accretions:<br />

293


Schiffer notes that burials, caches, and floors <strong>of</strong> structures may contain whole pots with<br />

remaining use life (Schiffer 1995c:178). Conversely, refuse deposits will seldom contain<br />

pristine articles. Postdepositional reworking can obscure remaining use life by breaking<br />

pots, but other types <strong>of</strong> use life evidence, such as the degree to which stone adzes have<br />

been re-sharpened in comparison to unused reject preforms, are more robust. Intuitively,<br />

ovenstones are a robust item that may preserve indications <strong>of</strong> remaining use life better than<br />

pottery, which could suggest site abandonment or lack <strong>of</strong> economising behaviour.<br />

Damage such as edge-rounding, surface ablation, spalling and chemical weathering<br />

<strong>of</strong> mineral constituents can give clues as to the use and taphonomic history <strong>of</strong> an artefact<br />

or fragment (see Schiffer 1995c:178-181 for a review). Water transport can cause rapid<br />

rounding <strong>of</strong> sherd edges (Skibo & Schiffer 1987) and abrasion by rolling can be<br />

distinguished (on lithic implements at least) from edge-rounding by smaller waterborne<br />

particles (Shackley 1974). Surface ablation and pedestalling <strong>of</strong> temper grains can result<br />

from water transport <strong>of</strong> sherds, and wetting/drying cycles in a saline environment can be<br />

expected to produce surface spalling or exfoliation (Schiffer 1987). In the extreme<br />

weathering environment <strong>of</strong> the tropical Pacific, calcareous temper grains are known to<br />

sometimes have dissolved after deposition, leaving voids in the fabric <strong>of</strong> sherds (see for<br />

example Dickinson 2000b), but there is no evidence for this occurring in sherds deposited<br />

in the sea. Sherds formerly on land and subsequently deposited in the sea as the result <strong>of</strong><br />

erosion might be expected to display this effect at some level, particularly if it is thought<br />

they were eroded from acid soils. Lack <strong>of</strong> weathering <strong>of</strong> calcareous temper grains may<br />

therefore be indicative in some settings <strong>of</strong> primary deposition in the sea.<br />

It was observed through differential conservation during the project that complete<br />

drying <strong>of</strong> sherds without desalination caused rapid degradation <strong>of</strong> the recovered<br />

assemblage, whereas even rough methods <strong>of</strong> desalination showed a marked improvement<br />

in the condition <strong>of</strong> analysed sherd surfaces and reduction <strong>of</strong> sherd crumbling. <strong>The</strong> absence<br />

<strong>of</strong> salt crystallization damage on sherds might suggest they have spent most <strong>of</strong> their lives<br />

294


in either a terrestrial or an underwater context, rather than in a cyclic wet/dry salty<br />

environment.<br />

Marine accretions have the potential to inform on taphonomic processes and<br />

landscape change. Sherds found on land that have accretion <strong>of</strong> marine organisms (coral<br />

growth for example) must have been in the sea at some time. Similarly, sherds with coral<br />

growth found in shallow water above the coral growth line may be evidence for landward<br />

transport, or for an emerging coastline. <strong>The</strong> converse is not true though, sherds found on<br />

land without such accretions need not have been deposited on land through their entire<br />

history. Abrasion and weathering can remove such accretions, which do not necessarily<br />

form in the first place.<br />

Artifact Diversity:<br />

<strong>The</strong> range <strong>of</strong> types <strong>of</strong> artifact in deposits is sensitive to cultural formation processes,<br />

especially the occupation span <strong>of</strong> settlements, but also to differences in the functions <strong>of</strong><br />

settlements and activity areas (Schiffer 1995c: 182-183). Schiffer’s discussion is weak on<br />

taphonomic effects on artefact diversity, and sampling effects on diversity generally (e.g.<br />

Kintigh 1984, Shott 1989a). Lag deposits, for example, may be size or density sorted,<br />

reducing diversity by favouring chunky types, those that break into large pieces, and<br />

different materials are differentially preserved in different depositional contexts, so there<br />

is always potential for taphonomic variation in diversity. Surface or lag deposits are more<br />

subject to collector effects, which can be expected to reduce diversity rapidly unless items<br />

are highly fragmented (Felgate & Bickler n.d). While Schiffer suggests that great diversity<br />

is found in secondary refuse deposits containing refuse streams from a settlement’s entire<br />

range <strong>of</strong> activities, this assumes a well-preserved deposit. Deposits lacking such diversity<br />

may do so as a result <strong>of</strong> taphonomic processes rather than cultural formation processes.<br />

Shott notes that richness as a measure <strong>of</strong> diversity, as used by Kintigh, fails to<br />

account for the evenness or equitability <strong>of</strong> item frequencies by class (Shott 1989a). Where<br />

295


large numbers <strong>of</strong> some classes <strong>of</strong> artefact, for example ceramics, exist at a location, and<br />

small numbers <strong>of</strong> another class, for example adzes, or where count data by class is sparse<br />

(low average abundance within classes), both richness and other more sophisticated<br />

measures <strong>of</strong> diversity such as the Shannon-Weaver information statistic can be sensitive<br />

to sample size effects. Shott goes on to suggest ways <strong>of</strong> inferring use-lives <strong>of</strong> tools from<br />

relative frequencies in diversity-saturated assemblages, like Schiffer, with minimal attention<br />

to the recognition <strong>of</strong> taphonomic reduction <strong>of</strong> diversity. He seems not to expect a strong<br />

taphonomic bias in the diversity <strong>of</strong> types within a single material class, e.g. lithics. It seems<br />

that a fundamental question in making use <strong>of</strong> diversity measures to identify behavioural or<br />

organizational (e.g. use-life) properties <strong>of</strong> assemblages is the effect <strong>of</strong> taphonomic biases<br />

on the data.<br />

Diversity therefore seems to me to be a poor route into the understanding <strong>of</strong><br />

formation processes for sites in a poor state <strong>of</strong> preservation, as indicated in the Roviana<br />

case (see Chapter 5).<br />

Representation <strong>of</strong> Parts:<br />

Use <strong>of</strong> part representation in ceramic assemblage formation studies is in its infancy<br />

(Schiffer 1995c:186) and preliminary ideas in this regard suffer from a convergence with<br />

inference from part representation in faunal taphonomy. Schiffer, for example, envisages<br />

part selection for recycling by potters as the sort <strong>of</strong> process that might be identifiable<br />

through part representation, rather like the way animal scavenging <strong>of</strong>ten involves selection<br />

<strong>of</strong> skeletal parts. Given that there is no evidence for use or re-use <strong>of</strong> sherds as grog temper<br />

in the Roviana materials, and only minor possible evidence for re-use <strong>of</strong> bases as bowls or<br />

frying pans (see Chapter 8), analysis along these lines would not seem to be a productive<br />

exercise.<br />

Other mechanisms, such as unrecorded collection by archaeologists and others, can<br />

be viewed as taphonomic processes affecting the composition <strong>of</strong> assemblages <strong>of</strong> artefacts<br />

296


(Ruig 1999). While there is as yet no generally accepted theory <strong>of</strong> the effects <strong>of</strong> collecting<br />

on ceramic part representation, some expectations are developed below in relation to the<br />

Roviana samples.<br />

Archaeologists and other destructive collectors are expected to pick up larger items<br />

first, and those with more information or aesthetic value. Whole adzes, more complete<br />

sherd or whole pots, are likely to be the first things collected from sites like the Roviana<br />

intertidal scatters. Next in order <strong>of</strong> preference are smaller items with obvious information<br />

content: decorated sherds, sherds on which the vessel lip, neck or shoulder/carination are<br />

represented, smaller lithic artefacts, or larger fragments <strong>of</strong> lithic artefacts. Some<br />

archaeologists may also recognize the information value <strong>of</strong> exotic lithic manuports such<br />

as sandstone fragments or ovenstones.<br />

Non-archaeologists might selectively remove sherds or lithics for purposes other<br />

than information gathering. Anecdotal evidence suggests two processes <strong>of</strong> this sort,<br />

skipping <strong>of</strong> sherds into deeper water by children, which should affect mainly sherds<br />

lacking complex curvature for hydrodynamic reasons, and collection <strong>of</strong> sherds and lithic<br />

artefacts by villagers following collection by archaeologists, where a perception <strong>of</strong><br />

commercial value has arisen. In this instance collector effects can be expected to closely<br />

mirror the information-value model <strong>of</strong> archaeological reconnaissance collection. A series<br />

<strong>of</strong> models <strong>of</strong> sample characteristics in line with these collection theories will be<br />

constructed below.<br />

Sediments:<br />

<strong>The</strong> sediments comprising the finer fraction <strong>of</strong> the matrix <strong>of</strong> sites may include<br />

archaeosediments (Waters 1992:16-17). Textural properties, mineralogy and chemical<br />

properties <strong>of</strong> sediments may inform on human behaviour (cultural formation processes) or<br />

on postdepositional process that have resulted in the formation <strong>of</strong> sites. For the Roviana<br />

sites, postdepositional sherd attrition as evidenced in Chapter 5 suggests the potential for<br />

297


a disaggregated pottery-derived component within site sediments. This is discussed further<br />

below in relation to models <strong>of</strong> site formation.<br />

This inventory <strong>of</strong> formation theory was used to develop a series <strong>of</strong> models <strong>of</strong><br />

formation and the material correlates <strong>of</strong> these. Models for cultural and natural formation<br />

processes are listed separately below, but there is substantial overlap in the correlates<br />

identified for the models, so data will be considered together, followed by a synthesis and<br />

conclusions.<br />

Models <strong>of</strong> Cultural Formation Process:<br />

<strong>The</strong> investigation <strong>of</strong> cultural formation processes is structured in terms <strong>of</strong> the following<br />

competing models:<br />

• Ceramic/lithic scatters formed as terrestrial discard from coastal settlements,<br />

subsequently inundated by the sea. <strong>The</strong> correlates <strong>of</strong> this model are evidence for<br />

coastal erosion features, poorly sorted artefact deposits on land and possibly sherd<br />

damage consistent with a terrestrial weathering regime, such as the weathering <strong>of</strong><br />

calcite grains.<br />

• Ceramic/lithic scatters formed as secondary accumulations in the sea when<br />

people living in terrestrial settlements dumped sherds and ovenstones into the<br />

sea. <strong>The</strong> correlates <strong>of</strong> this model are evidence for structures and midden on land<br />

from former settlements, or erosional features that suggest obliteration <strong>of</strong> such<br />

evidence.<br />

• Ceramic/lithic scatters are discard from stilt settlements in the<br />

intertidal/subtidal zone. Expected correlates are minimal contemporaneous<br />

evidence for occupation on adjacent land, with any ceramic-lithic evidence on land<br />

showing evidence <strong>of</strong> size or density sorting, and associated with geomorphological<br />

298


stormwash features; we would not expect to see much in the way <strong>of</strong> stone or coral<br />

building materials scattered about if timber posts were the primary means <strong>of</strong><br />

supporting post and thatch structures.<br />

• Ceramic/lithic scatters formed as discard from now-defluved artificial islet<br />

settlement located in the intertidal. This should look similar to stilt houses, and<br />

we might also expect to see a scatter <strong>of</strong> coral stones or boulders that formerly<br />

comprised the sea-walls <strong>of</strong> such structures, although mining for building materials<br />

in later periods might remove such evidence.<br />

Some additional cultural formation models that are not mutually exclusive with the above<br />

list can be advanced:<br />

A. Site abandonment might have a correlate <strong>of</strong> remaining use life <strong>of</strong> adzes and<br />

ovenstones, where dicarded adzes show little sign <strong>of</strong> resharpening, and many large<br />

unfractured ovenstones remain on site, but mining <strong>of</strong> ovenstones and collector<br />

effect on adzes could remove such evidence.<br />

B. Whether terrestrial or over water, permanent settlement should have the dual<br />

correlates <strong>of</strong> diversity <strong>of</strong> artefact types and substantial breakage populations <strong>of</strong> a<br />

diverse range <strong>of</strong> functional types <strong>of</strong> pottery; conversely, short-term and/or low-<br />

intensity use should be evidenced by smaller accumulations and variable artefact<br />

inventories.<br />

C. Specialized site functions should result in low-diversity artifact assemblages, with<br />

preponderance <strong>of</strong> particular vessel forms. Conversely, generalized settlement<br />

should have a diversity <strong>of</strong> special-purpose forms alongside general-purpose forms.<br />

Models <strong>of</strong> Post-depositional Formation Process:<br />

299


Substantial sherd attrition through post-discard processes (as shown to have occurred in<br />

the Roviana cases in Chapter 5) raises the prospect <strong>of</strong> taphonomic sampling bias in the<br />

surviving sherd population. In addition to the cultural formation models outlined in the<br />

previous section, a series <strong>of</strong> models <strong>of</strong> sherd attrition factors will be constructed and<br />

considered in relation to the Roviana data. <strong>The</strong>se, with their correlates, are outlined below.<br />

A. <strong>The</strong> transport model, where the missing sherds were transported by wave action<br />

into the higher-energy environment <strong>of</strong> the strandline and destroyed by mechanical<br />

abrasion. Under this model sherds which were sheltered from wave energy by local<br />

microtopography or sediment armouring form either a random selection <strong>of</strong> sherds,<br />

or a selection structured by sherd transportability. Also, weak sherds should be<br />

recovered, but extremely thin or light sherds should exist principally as small<br />

fragments on the strandline. Sites with a greater wave exposure should show<br />

greater evidence for size-sorting.<br />

B. <strong>The</strong> dis-aggregation model, where poorly-fired sherds or sherds otherwise having<br />

poor fabric strength have disintegrated in place regardless <strong>of</strong> local<br />

microtopography. Under this model, s<strong>of</strong>t or weak sherds should not be present in<br />

assemblages, although porous sherds suffering salt damage (Schiffer 1987:277) on<br />

retrieval may be present. An assessment <strong>of</strong> fabric strength variability (as expressed<br />

by sherd size) in relation to temper types <strong>of</strong> the recovered sherds will be presented<br />

in this <strong>chapter</strong> to try to determine whether much strength variability has survived.<br />

If it has, this favours a transportation model, and if not, a disaggregation model is<br />

favoured. Sherd size in relation to thickness and fabric will be used as a proxy for<br />

sherd strength. No experimental strength tests on sherds were performed. (As will<br />

be argued below in relation to signs <strong>of</strong> weathering, post-depositional reduction in<br />

sherd strength does not seem to have occurred in any <strong>of</strong> the fabric groups during<br />

their time in the sea.)<br />

300


C. <strong>The</strong> collector model, where the ratio <strong>of</strong> plain to decorated diagnostic sherds is<br />

substantially different between sites. One assemblage is known to have been<br />

depleted in the past through collection by archaeologists and villagers, and this site<br />

will be used as a control in identifying the effects <strong>of</strong> sherd collection by humans.<br />

<strong>The</strong> number <strong>of</strong> sherds for which vessel part representation is high can be expected<br />

to vary with assemblage brokenness, but also with the level <strong>of</strong> collecting by humans<br />

in the past. Assemblage brokenness can be compared using body sherd size as a<br />

control, and the ratio <strong>of</strong> eye-catching diagnostic sherds to body sherds can be<br />

developed as an indicator <strong>of</strong> collecting effects in the past. One might however<br />

expect some larger body sherds to be uplifted, which might render this model<br />

equifinal with brokenness through wave exposure or trampling.<br />

D. <strong>The</strong> sherd-skipping model, where anecdotal evidence for sherd skipping by<br />

children (throwing a sherd or flat stone with a spinning action across the water<br />

surface to cause it to perform a series <strong>of</strong> bounces) is tested against assemblage<br />

composition. Assemblages that have been subject to a high degree <strong>of</strong> sherd<br />

skipping should have relatively fewer flattish medium-sized body sherds than would<br />

otherwise be the case.<br />

E. <strong>The</strong> collection intensity model: in this model sherd size varies in relation to<br />

collection intensity. <strong>The</strong> more intensively a site is collected, the more small<br />

inconspicuous or undiagnostic fragments are retained into the sample assemblage.<br />

Summary <strong>of</strong> Models:<br />

<strong>The</strong> consequences <strong>of</strong> the actions <strong>of</strong> these various processes for archaeological sequence-<br />

building objectives, as laid out at the beginning <strong>of</strong> the <strong>chapter</strong>, vary: the transport model<br />

is liable to create bias in favour <strong>of</strong> heavier, thicker pottery, with light calcareous fabrics<br />

301


possibly under-represented, and thin-walled pottery particularly likely to be under-<br />

represented. <strong>The</strong> dis-aggregation model is likely to result in the under-representation <strong>of</strong><br />

non-cooking vessels, which may be lower-fired, and/or less tough through choice <strong>of</strong><br />

temper during manufacture.<br />

This possibility is <strong>of</strong> particular concern for Lapita, since some have suggested that<br />

highly decorated flat-based unrestricted vessels might have a serving function, or that<br />

highly decorated Lapita vessels in general might have a ritual function rather than<br />

utilitarian. If this were the case a harsh taphonomic regime might turn Lapita sites into<br />

post-Lapita sites from the point <strong>of</strong> view <strong>of</strong> the ceramic chronologist. Wickler’s<br />

documentation <strong>of</strong> dentate-stamped pottery in sites <strong>of</strong> this general type on the reef flat at<br />

DAF and DES are reassuring in this regard, suggesting that there is a possibility <strong>of</strong> such<br />

sherds being represented in un-buried intertidal/shallow-water sites (Wickler 2001:Chapter<br />

5). Collector bias might render an assemblage plainer overall than it might have been, but<br />

degradation <strong>of</strong> the overall sample <strong>of</strong> diagnostic sherds rather than analytical bias might be<br />

expected, as the analyst can control for this effect by controlling for sherd<br />

“diagnosticness”. <strong>The</strong> sherd skipping model does not seem to involve any particular<br />

stylistic taphonomic bias, except if decoration is concentrated on the upper portions <strong>of</strong> the<br />

potsherd skipping could affect the overall ratio <strong>of</strong> plain:decorated sherds, but many other<br />

factors could also affect this.<br />

Qualitative Evidence From Samples:<br />

Damage:<br />

Wetting/drying salt crystallization damage was not quantified during analysis, but sherd<br />

condition and the spatial distribution <strong>of</strong> sherd condition was used to infer drying regime<br />

limits for major portions <strong>of</strong> sites in the past. Most sherds at the deeper margins <strong>of</strong> sites<br />

exhibit no sign <strong>of</strong> salt erosion damage or damage arising as a result <strong>of</strong> water transport<br />

302


(surface ablation, pedestalling <strong>of</strong> temper grains, and severe edge rounding). It was inferred<br />

that these have never been exposed to severe, sustained swash zone turbulence (rolling)<br />

or drying. Also, there was no sign <strong>of</strong> postdepositional weathering <strong>of</strong> carbonate temper<br />

grains suggesting the sherds had never been exposed to an acid environment. Some<br />

carbonate grains were partially sintered, with a rind <strong>of</strong> carbonate still lining the sides <strong>of</strong> the<br />

void around a shrunken grain, thought to be an artefact <strong>of</strong> manufacturing firing<br />

temperature, but many sherds had un-sintered carbonate grains which aught to be liable to<br />

solution if the sherds had been in an acid environment. Most sherds had slight edge<br />

rounding, consistent with abrasion by waterborne sand.<br />

Accretions:<br />

No quantitative analysis was conducted, but some general statements are made. Accretions<br />

<strong>of</strong> marine growth were almost universal, except for rolled sherds found on the strandline,<br />

and at the Gharanga stream-mouth site where salinity was probably low. Accretions<br />

included dead hard coral in some cases, live sponges and coralline/green algae in most<br />

cases. In the deeper margins <strong>of</strong> some sites dead coral growth was common on sherds<br />

(Honiavasa site for example) in spite <strong>of</strong> the lack <strong>of</strong> any evidence for active coral growth<br />

nowadays. <strong>The</strong>se accretions may be evidence for a high stand at or post deposition, which<br />

is consistent with current data on sea-level history (Mann et al. 1998). Dating <strong>of</strong> accretions<br />

might be a worthwhile future exercise for palaeo-sea-level studies.<br />

Had wave action transported sherds up the shore in significant quantity one would<br />

expect to find coral accretions on land as evidence. It appeared as though there was a<br />

significant correlation between sherd condition and accretion presence; sherds from deeper<br />

water tended to have coral and sponge accretions and were also larger and in better<br />

condition. No coral accretions were seen on land, but this may be a sample effect, as the<br />

number <strong>of</strong> sherds located on adjacent land was small (one sherd from Honiavasa, two or<br />

three small sherds from Gharanga, none from Hoghoi). It could also be an abrasion effect.<br />

Sherds were seen in the roots <strong>of</strong> a fallen tree a short distance inland from the Honiavasa<br />

site but extensive testpitting in the area did not reveal substantial pottery deposits<br />

(Sheppard, pers. com.).<br />

Artifact Diversity:<br />

Artefact diversity/inventory <strong>of</strong> material is discussed in Chapter 13.<br />

303


Sediments:<br />

In regard to the model <strong>of</strong> cultural formation <strong>of</strong> settlements located on artificial islets in the<br />

past, now reworked, some sites were characterized by a scatter <strong>of</strong> easily transportable<br />

coral blocks <strong>of</strong> the type used locally these days for building wharves and islets. Although<br />

no quantified data were collected that bear directly on this issue, it was noted during<br />

fieldwork that the gravels and coral cobbles <strong>of</strong> some site sediments could be the reworked<br />

remains <strong>of</strong> such structures (especially at Hoghoi). No specific patterned features were<br />

noted. As it is, only the speculative suggestion can be made, and this is a question that<br />

must be left to future research. Cultural formation processes, like sea levels, should not be<br />

expected to have been static over time, and the changing conformation <strong>of</strong> the shoreline<br />

pr<strong>of</strong>ile with falling sea-levels may have enabled or encouraged such cobble structures later<br />

in time, where earlier the depth <strong>of</strong> water over the current reef flat may have precluded islet<br />

construction as too labour intensive.<br />

Micro-examination <strong>of</strong> two sediment samples from Hoghoi suggest that substantial<br />

quantities <strong>of</strong> ferromagnesisn opaque minerals and clinopyroxene were present in one <strong>of</strong> the<br />

samples, and minor quantities <strong>of</strong> fine ferromagnesian minerals and occasional small lathes<br />

<strong>of</strong> pyroxene were present in the other. <strong>The</strong> first sample was collected by the author from<br />

the vicinity <strong>of</strong> the densest accumulations <strong>of</strong> pottery at the site, while the second was<br />

collected from an unknown location in the general vicinity <strong>of</strong> the site (Sheppard pers com).<br />

<strong>The</strong>se results have yet to be confirmed or quantified by thin-section petrography.<br />

Remaining Use Life:<br />

All pottery seen by the team was broken, although anecdotes <strong>of</strong> “whole” pots in deeper<br />

water were heard. Two complete adzes and a number <strong>of</strong> fragments were recovered, and<br />

an unground adze preform in the possession <strong>of</strong> a local landholder was photographed (See<br />

Chapter 10). Comparison <strong>of</strong> the recovered whole adzes and the preform suggested that<br />

both whole adzes found had been heavily re-sharpened, and were too short to have much<br />

remaining use life in their intended function. Ovenstones were systematically collected at<br />

Hoghoi. Many large fragments and some complete water-rounded cobbles were found,<br />

enough to use for cooking, suggesting remaining use life existed for these. <strong>The</strong> presence<br />

<strong>of</strong> these may have resulted from abandonment, or, more likely in view <strong>of</strong> the lack <strong>of</strong> adze<br />

remaining use life, the presence <strong>of</strong> these at Hoghoi may simply indicate a lack <strong>of</strong><br />

economising behaviour, or perhaps discard <strong>of</strong> insufficiently tough stones liable to explode.<br />

<strong>The</strong> stones may have had other, non-cooking functions such as canoe or net anchors.<br />

304


Quantitative Evidence: Size/Density Effects and the Structure and Context <strong>of</strong> the<br />

Deposit:<br />

Sherd Size and Sherd Thickness (Thickness Strength):<br />

To investigate the relationship between thickness and sherd size, while controlling for<br />

fabric variation and form variation, body sherds <strong>of</strong> the most numerous temper class<br />

(placered volcanic temper) were used. It was initially expected that strength, and thus area<br />

<strong>of</strong> recovered sherds would show a positive correlation with thickness, which proved to be<br />

the case (Figure 76).<br />

Control for gross variation in form strength was achieved by selecting out only<br />

those sherds without complex form, that is, body sherds. This means that many <strong>of</strong> the<br />

larger sherds were omitted, due to representation <strong>of</strong> more complex forms likely to confer<br />

sherd strength. Those too small to assign to vessel part were also omitted. Sherd area<br />

seems to plateau above 9mm thickness, with the greater area <strong>of</strong> the 13mm and 15mm<br />

categories possibly being a sample size effect. Thin sherds (less than 8mm thick) decline<br />

in median size roughly proportional to thickness, and sherds less than 4mm thick are rare.<br />

Sherds in the 4-8mm thickness range form the bulk <strong>of</strong> the total sample, while sherds in the<br />

8-12mm thickness range yield consistently larger median sherd sizes. Thickness greater<br />

than 12mm is rare, and seems to yield more variable sherd size, although this might be a<br />

function <strong>of</strong> small sample size. <strong>The</strong>se data seems to support the hypothesis that there may<br />

be a bias against very thin wares in the recovered sample, which are liable to exist only as<br />

small sherds fortuitously preserved rather than transported to the strandline and rolled to<br />

destruction. Large thin sherds are either overly fragile and became broken into small, easily<br />

transported thin sherds, or large thin sherds are more easily transported by water in the first<br />

place, and thus tended to be destroyed by transport to and abrasion in the swash zone.<br />

Temper Class and Sherd Size:<br />

Temper classes are detailed in Chapter 4. Without controlling for thickness, but looking<br />

at just body sherds, sherd size seems only weakly related to temper class (Figure 77),<br />

except in the cases <strong>of</strong> Groups 4 and 7, where sample size was inadequate. Group 5<br />

(Quartz-calcite) median size is slightly smaller.<br />

305


Figure 76: Effect <strong>of</strong> sherd thickness on sherd strength<br />

(controlled for temper variation by using placered volcanic<br />

tempered sherds only).<br />

Figure 77: Body sherd size for the various temper classes.<br />

306


Body Sherd Size Frequency by Site:<br />

<strong>The</strong> sherd skipping model predicts depletion in mid-sized (throw-sized,10-30cm 2 ) body<br />

sherds, but utility <strong>of</strong> this model was limited by the potential for assemblage brokenness to<br />

be equi-finally structured by other postdepositional factors. Analysis was limited to those<br />

sites which retain a reasonable corpus in their collected assemblages <strong>of</strong> body sherds larger<br />

than 30cm 2 . <strong>The</strong> dataset provided 1053 sherds comprising body part <strong>of</strong> the vessel only,<br />

from eight sites, but insufficient larger sherds remained to test the sherd skipping model<br />

using this method. Body sherds in the 20cm 2 size class predominate in the Miho site<br />

(Figure 78), located at a modern village from where anecdotes <strong>of</strong> sherd skipping were<br />

heard, and at Honiavasa, across a narrow channel from the same village, so either there is<br />

something wrong with my theory <strong>of</strong> the correlates <strong>of</strong> this behaviour, or other processes<br />

have structured the assemblages to a greater degree.<br />

Differences between sites in these histograms are largely attributed to collection<br />

intensity: for some sites sherds in the smallest size category (4cm 2 ) are numerous, while<br />

absent in others, a pattern which dominates variability between assemblages. Low-intensity<br />

collections like Honiavasa, Gharanga and Miho have few sherds in the 4mm size class,<br />

while these predominate in the intensively collected sites Zangana, Paniavile and Hoghoi<br />

This is clearly not the whole story though, as despite low collection intensity Honiavasa<br />

yielded many more large (>56cm 2 ) sherds than either Hoghoi, Miho, Gharanga or Zangana.<br />

Representation <strong>of</strong> Parts in the Site Samples:<br />

<strong>The</strong> collector model predicts that assemblages previously collected by archaeologists or<br />

other interested people will be depleted <strong>of</strong> conspicuous large or highly decorated sherds,<br />

particularly those with information pertaining to upper vessel form (lip, rim, neck and<br />

shoulder). Lip/Rim-only sherds and body-only sherds were counted by site. <strong>The</strong>se counts<br />

together with the ratios between them may provide an enumerator <strong>of</strong> previous collection<br />

intensity, operating on the assumption that body sherds have a greater probability <strong>of</strong> being<br />

ignored by a collector (Table 23). Only sherds greater than 15cm 2 were selected, to<br />

control to some extent for assemblage brokenness.<br />

307


Figure 78: Histogram <strong>of</strong> body sherd size classes by site.<br />

308


Table 23: Ratio <strong>of</strong> lip-rim sherds to body sherds as a measure <strong>of</strong> collector effect.<br />

SITE Total >15cm 2 Lip+Rim<br />

count >15cm 2<br />

309<br />

Body count<br />

>15cm 2<br />

Paniavile 184 17 71 0.239<br />

Hoghoi 155 31 66 0.47<br />

Miho 205 39 58 0.672<br />

Honiavasa 319 75 70 1.071<br />

Gharanga 135 20 41 0.479<br />

Nusa Roviana 40 5 9 0.555<br />

Kopo 22 3 10 0.3<br />

Zangana 222 37 56 0.661<br />

ratio <strong>of</strong> LR:B<br />

It seems the ratio <strong>of</strong> lip-rim sherds larger than 15cm 2 to body sherds larger than 10cm 2 may<br />

well be a good measure <strong>of</strong> collector effect, as the most collected site (prior to the present<br />

study) was Paniavile (previously yielding 84 rim sherds, 41 “decorated body sherds” and<br />

an uncounted number <strong>of</strong> plain body sherds in an initial collection event reported by Reeve<br />

1989:50).<br />

What then <strong>of</strong> the other sites? Nusa Roviana and Kopo yielded small samples <strong>of</strong><br />

sherds, and their “collector” ratios are best ignored. Hoghoi and Gharanga both have ratios<br />

nearing 0.5, which provide a weak suggestion that some <strong>of</strong> the more complete sherds may<br />

have been removed from these in the past, prior to initial collection in the course <strong>of</strong> the<br />

present study. Miho and Zangana both yield ratios approaching 0.7, suggesting this is a<br />

reasonable figure for newly-discovered sites collected in a moderately intense fashion (the<br />

bulk <strong>of</strong> these collections being obtained in single days by two and five collectors<br />

respectively). Honiavasa, collected by a lone archaeologist on a rising tide over about four<br />

hours, looks most like a first-hit, low intensity collection <strong>of</strong> large complex-shape sherds,<br />

analogous, perhaps, to the original Reeve/Spriggs Paniavile surface collection in vessel


part representation. <strong>The</strong> major difference between the Paniavile ratio (0.24) and that <strong>of</strong><br />

the most similar sample (Hoghoi, 0.47), reassures to some extent that the major collections<br />

other than Paniavile are not severely biased by previous unreported collection, and alerts<br />

us to the degraded information content <strong>of</strong> the Paniavile sample as used in the current<br />

research.<br />

Frequency <strong>of</strong> Decoration:<br />

This is potentially another way <strong>of</strong> measuring collector effects. While variation in<br />

production in the past might account for different frequency <strong>of</strong> decoration in assemblages,<br />

collector effects could create the same result, so this is something to be aware <strong>of</strong> in<br />

analyzing surface collections. It might be simplest to avoid using frequency <strong>of</strong> decoration<br />

in seriations etc., but as others do this (see Chapters 1&2) it was thought to be worth a<br />

short detour to investigate this topic. For each vessel part, the number <strong>of</strong> plain instances<br />

and the number <strong>of</strong> instances <strong>of</strong> decoration were counted. <strong>The</strong> ratio <strong>of</strong> undecorated to<br />

decorated was calculated (Table 24).<br />

For all vessel parts except shoulders the previously collected Paniavile site scored<br />

within the range <strong>of</strong> the other sites. <strong>The</strong>se data are interpreted as showing that the ratio <strong>of</strong><br />

decorated to undecorated sherds can remain stable through collection events. It should be<br />

Table 24: Ratios <strong>of</strong> plain to decorated sherds, controlled by vessel part (lips, rims,<br />

necks and shoulders.<br />

SITE Plain L:Dec L Plain R:Dec R Plain N:Dec N Plain S:Dec S<br />

1 (Paniavile) 20:59 0.34 110:36 3.1 98:31 3.2 56:26 2.2<br />

2 (Hoghoi) 34:51 0.64 113:25 4.5 134:18 7.4 121:22 5.5<br />

3 (Miho) 14:52 0.24 80:35 2.3 77:30 2.6 84:14 6.0<br />

4 (Honiavasa) 23:80 0.29 154:17 9.1 153:20 7.7 125:23 5.4<br />

5 (Gharanga) 10:24 0.42 36:15 2.4 62:12 5.7 49:21 2.3<br />

8 (Zangana) 36:44 0.82 95:39 2.4 137:36 3.8 114:30 3.8<br />

310


noted in this regard that initial surface collection <strong>of</strong> a full cover sample <strong>of</strong> sherds from<br />

Zangana targeted only decorated sherds, with subsequent 1/5 and full cover collections<br />

targeting all sherds and large sherds respectively. Zangana may thus be slightly enriched<br />

decoratively through the initial collection strategy, but this effect is not large.<br />

Other patterns worthy <strong>of</strong> note in Table 24 are that the Honiavasa site gains a marginal<br />

score in two out <strong>of</strong> four vessel part categories (rim and neck), while lips are more likely<br />

to be decorated than in other sites and rims and necks are more likely to be plain, showing<br />

that overall ratio <strong>of</strong> decorated to plain sherds, without control <strong>of</strong> location on the vessel,<br />

can be something <strong>of</strong> a lottery rather than a salient measure <strong>of</strong> assemblage decorative<br />

variability. Miho, also, for example, has the most decorated lip, rim and neck scores, but<br />

the plainest shoulders. Breakdowns <strong>of</strong> the specifics <strong>of</strong> decoration contributing to these<br />

broad categories <strong>of</strong> data are given in detail in Chapter 9.<br />

Another index <strong>of</strong> sherd skipping was considered, the ratio <strong>of</strong> neck sherds (not a<br />

good skipping shape from personal experience) to body sherds, the latter <strong>of</strong> size range<br />

>12cm 2 ,


epresented would be too sensitive to assemblage brokenness. This index is problematic<br />

as it may relate to overall assemblage vessel form, and/or to collection bias rather than to<br />

postdepositional sherd skipping. <strong>The</strong>se data are presented in Table 25.<br />

<strong>The</strong> weakest <strong>of</strong> these “Skipping indices” is for the Paniavile site, heavily collected<br />

prior to the present study, suggesting again that previous collection may have been biased<br />

against body sherds. <strong>The</strong> lowest (strongest) index was that <strong>of</strong> Honiavasa, despite a<br />

tendency for sherds to break at the neck in this site, probably as a result <strong>of</strong> slab<br />

construction. Zangana is not far separated, suggesting Zangana, at least, might have been<br />

subject to selective removal <strong>of</strong> mid-sized body sherds, particularly since, the intensity <strong>of</strong><br />

collection at Zangana was quite high. An alternative, and more likely possibility is that the<br />

high neck count at Zangana relates to an overall pattern in vessel form where the necked<br />

vessel is more common in production, or more robust and more commonly preserved.<br />

Miho is next strongest in this index, and as Miho is located at the anecdotal skipping<br />

village, and Honiavasa within a short distance, there may be something to it.<br />

Sea Level History:<br />

Radiocarbon ages <strong>of</strong> uplifted corals, combined with an assumed high stand <strong>of</strong> +1m at<br />

5500BP generate mean uplift rates for the New Georgia tectonic block <strong>of</strong> between<br />

0.2mm/yr and 0.9mm/yr, with the three data locations in Roviana Lagoon area suggesting<br />

mean tectonic uplift rates <strong>of</strong> 0.7mm per year (Rereghana Island and Araroso passage) to<br />

0.9mm per year (Nusa Roviana Island) (Mann et al. 1998: Figure 8, Table 1, Table 2.). If<br />

sea levels have been falling an average <strong>of</strong> 0.2mm per year since the late-Holocene high<br />

stand, and the land has being going up at an average 0.7 to 0.9 mm per year in the Roviana<br />

area, then the combined effect over the 2,800 years may be somewhere in the vicinity <strong>of</strong><br />

an effective fall in sea level <strong>of</strong> between 2.5 and 3.0 metres. It is unlikely that anything in<br />

312


nature happens this smoothly, but there is currently no local data for the period between<br />

about 6000BP and 200BP, leaving little alternative but to extrapolate. Data from Ruvi Bay<br />

on Kolombangara is reassuring in this respect, as coral dated there to 2250BP ±60 is one<br />

metre above sea level, and as uplift rates there seem substantially lower (0.1 mm/yr,<br />

calculated from a total Holocene emergence <strong>of</strong> 1.5m) suggesting that if a correction is<br />

applied for the greater uplift at Roviana Lagoon <strong>of</strong> 0.6-0.8mm per year, we could add<br />

1.35-1.8 metres on to this (if the average uplift rates are useful) and that Roviana sea level<br />

(relative) might have been about 2.35-2.8m higher up the shoreline at around 2250BP.<br />

More data from the Roviana area is needed before this figure can be accepted with any<br />

certainty. All paleo-shoreline evidence, however, points to an emergent coastline, with the<br />

artefact scatters in deeper water (possibly buried in s<strong>of</strong>t lagoonal sediments) until fairly<br />

recently. This is consistent with cultural formation processes for the artefact scatters as<br />

discard from anchored vessels, fishing platforms, or more substantial stilt houses. This<br />

conclusion would need to be revised, though, should the gradual sea level model used to<br />

extrapolate between insufficient data points turn out to be incorrect.<br />

Chapter Summary and Conclusions:<br />

This <strong>chapter</strong> was written from the perspective suggested by the results <strong>of</strong> the analysis in<br />

Chapter 5 that most <strong>of</strong> the sherdage discarded in the past has not made it into our sample,<br />

and <strong>of</strong> the missing sherds, most are no longer extant. It has not been possible to provide<br />

a definitive account <strong>of</strong> where the missing pottery went, but a brief synthesis <strong>of</strong> the<br />

analytical results in hand is given below, to clarify which <strong>of</strong> the models set up at the outset<br />

<strong>of</strong> this <strong>chapter</strong> are consistent with the properties <strong>of</strong> the collections.<br />

Cultural Formation Processes:<br />

<strong>The</strong> absence <strong>of</strong> adjacent settlement evidence on land for all intertidal sites argues against<br />

313


models <strong>of</strong> terrestrial settlement. <strong>The</strong> good state <strong>of</strong> preservation <strong>of</strong> Acropora corals on the<br />

adjacent shore platform, and the evidence from sea level studies (Mann et al. 1998)<br />

suggest that this feature is an emerged reef flat which would have been the upper limit <strong>of</strong><br />

coral growth at the time the sites were formed. Any ceramic deposits on this flat would be<br />

highly visible, and would have been deposited into the sea in the Lapita period. <strong>The</strong>se<br />

factors strongly support the conclusion <strong>of</strong> a maritime rather than terrestrial origin.<br />

Additional evidence against terrestrial models is the complete absence <strong>of</strong><br />

weathering <strong>of</strong> carbonate grains, which sometimes occurs in terrestrial deposits in the<br />

tropical Pacific. Reefal carbonate detritus used as tempering material is invariably in<br />

pristine condition except where slight sintering has occurred during firing.<br />

Accretions provide an additional suggestion <strong>of</strong> a high stand <strong>of</strong> sea level at-or-post-<br />

deposition. Lack <strong>of</strong> salt drying damage in the deeper margins <strong>of</strong> sites, and severe damage<br />

<strong>of</strong> this sort to many sherds post-recovery, suggest that gradual submergence <strong>of</strong> terrestrial<br />

sites is unlikely to account for the presence <strong>of</strong> the scatters in the sea.<br />

Some sites could conceivably have been located on piled coral wharves or artificial<br />

islets, now eroded or otherwise reworked. <strong>The</strong> shoreline and intertidal at Hoghoi has many<br />

coral blocks <strong>of</strong> the size used in this sort <strong>of</strong> structure today, but these are common along<br />

extensive stretches <strong>of</strong> the raised-reef barrier island where Hoghoi is located, even where<br />

there is no pottery. Better-preserved buried sites from the same general period in the<br />

Bismarck archipelago (Gosden 1989, 1991b, Gosden et al. 1989, Gosden & Webb 1994,<br />

Kirch 2001), seemed to have post structures, and there is no strong evidence yet for any<br />

different pattern at Roviana Lagoon.<br />

Postdepositional processes are likely to have removed any ceramic evidence for site<br />

abandonment. While there was some remaining use life in “ovenstones” collected from the<br />

Hoghoi site, the two complete adzes recovered were heavily re-sharpened if the<br />

proportions <strong>of</strong> an unused reject preform in a local collection were anything to go by. <strong>The</strong><br />

ovenstones could indicate lack <strong>of</strong> economizing behaviour rather than abandonment.<br />

314


Post-deposition Taphonomic/Formation Processes:<br />

In the disaggregation model outlined above, it was predicted that sherd strength variability<br />

would be low. Analysis <strong>of</strong> sherd size in relation to temper suggests that strength variability<br />

between temper types as reflected by sherd size is low for the assemblages. Whether this<br />

is due to congruence in technical competence using differing materials, or whether the<br />

weaker sherds have already gone is uncertain. Favouring the latter view was the presence<br />

among the stronger sherds <strong>of</strong> infrequent s<strong>of</strong>t sherds or (very rarely) partially dissolved<br />

sherds with solution scarring, suggesting that disaggregation has been a factor in sherd<br />

attrition. Additional support for a disaggregation model was found when sediments from<br />

the Hoghoi site were examined microscopically, and found on examination under low-<br />

power magnification to contain abundant volcanic mineral grains exotic to the calcite<br />

raised coral barrier island local geology. <strong>The</strong>se grains are thought to be temper sands<br />

deriving from disaggregated pottery, but this conclusion has yet to be confirmed<br />

petrographically.<br />

<strong>The</strong> data on sherd thickness indicates that thinner sherds are weaker (break more<br />

easily) which makes the surviving fragments more easily transported by wave action, and<br />

possibly also that large thin sherds have been transported to the strandline and destroyed<br />

(weak fabrics and large thin shapes are selected against by a transport model). This model<br />

was consistent with the frequency distribution by thickness in which small thin sherds were<br />

rare. Large, thick sherds were rare also, which could relate to behavioural factors such as<br />

use life, but may also have resulted from inadequate firing <strong>of</strong> thicker sherds, and thus less<br />

resistance to breakage/transport and to disaggregation.<br />

<strong>The</strong>re is some support in the part representation data for the collector model and<br />

the collection intensity model. <strong>The</strong> collector model is supported by the ratio <strong>of</strong> lip/rim<br />

sherds to body sherds, which showed Paniavile to have been more heavily collected in the<br />

past than other sites (corresponding with events reported in Reeve 1989). Body sherd size<br />

histograms showed correlation with collection intensity for assemblages in the present<br />

study.<br />

315


Sherd skipping behaviour has not left any obvious hole in histograms <strong>of</strong> body sherd<br />

size-frequency, although there was some support for the model in the ratio <strong>of</strong> body sherds<br />

to neck sherds, but underwater archaeology might be a better way to ascertain the extent<br />

<strong>of</strong> recent sherd skipping or sherd throwing into deeper water.<br />

Interpretation:<br />

It seems likely that different processes have predominated at different times. During the<br />

period <strong>of</strong> occupation, stilt houses would have been constructed in water possibly too deep<br />

to stand in at low tide, with sharpened posts driven into lagoonal mud until they reached<br />

the coral underneath. At this stage there was probably a zone <strong>of</strong> shallow-water Acropora<br />

corals to landward <strong>of</strong> some sites, backed by a low cliffed shoreline (Hoghoi, Miho,<br />

Zangana, and Paniavile at least, although this is not immediately apparent at Zangana today<br />

due to landscape alterations during WWII, and possibly Nusa Roviana also), making canoe<br />

landing and access difficult. Early in the history <strong>of</strong> stilt settlements, the pottery is likely to<br />

have been semi-buried in s<strong>of</strong>t lagoon sediments, with disaggregation due to dissolution,<br />

and turbation by plants, grazing animals and burrowing animals and<br />

construction/maintenance being major factors contributing large quantities <strong>of</strong> exotic sands<br />

to the local sediments. Falling sea levels post-abandonment are likely to have brought the<br />

ceramics into the lower margins <strong>of</strong> the swash zone fairly recently, removing organics and<br />

fine sediments to a varying extent depending on the site location, and leaving lag deposits<br />

<strong>of</strong> stronger sherds. Wave transport and anthropogenic processes such as sherd skipping<br />

may have caused gradual or intermittent sherd attrition to a greater extent than other<br />

mechanisms in more recent years. It seems from the Rereghana oyster evidence presented<br />

by Mann et al. that the current sea levels are very recent, and have been at their present<br />

level no more than 400 years, so these latter processes are unlikely to have affected the<br />

sites over the millennia. In the modern era, particularly over the last three decades, interest<br />

in the sites by archaeologists and others has had a major impact on one site (Paniavile),<br />

selectively removing large, decorated, and rim sherds initially, with more intensive<br />

316


collection <strong>of</strong> a number <strong>of</strong> sites during the present research.<br />

Some <strong>of</strong> the most significant taphonomic biases may have occurred during the early<br />

disaggregation phase <strong>of</strong> sherd attrition, where particular functional variants or production<br />

groups may have disappeared almost entirely, due to low firing or incomplete firing <strong>of</strong> the<br />

vessel wall interior. <strong>The</strong>se types <strong>of</strong> biases would have to be taken into account in<br />

comparing intertidal sites with terrestrial sites, as overall effect on assemblage composition<br />

could be large. Evidence will be presented elsewhere (Chapters 8 and 9) that at least one<br />

production style had a fragile tall, thin excurvate lip, and that as a result the style is known<br />

largely through rim/neck/shoulder sherds rather than lip/rim/neck/shoulder sherds.<br />

Transport processes are likely to have biassed preservation towards chunkier types which<br />

broke into large pieces, like Honiavasa carinated sherds, and biases between intertidal sites<br />

can be expected due to slight differences in wave exposure.<br />

Of the taphonomic models constructed in the introduction, it is the collector and<br />

collection intensity models that are most clearly supported by the size data and the part<br />

representation data in this <strong>chapter</strong>. Collector processes seem to have impacted severely on<br />

the information content <strong>of</strong> sites, leaving the Paniavile sample stylistically depauperate in<br />

information on vessel form and decoration.<br />

A methodological conclusion arising from this result is that reconnaissance survey<br />

and recording <strong>of</strong> surface intertidal sites in the future could be formulated in the light <strong>of</strong><br />

these findings, i.e. that the effects <strong>of</strong> casual unprovenanced surface collections on the<br />

information content <strong>of</strong> intertidal surface scatters is drastic. Where possible, in-situ non-<br />

destructive recording is advocated unless detailed spatial recording (point provenancing<br />

<strong>of</strong> all items, including accurate heights in relation to a datum) and field materials<br />

conservation can be carried out. One potential problem with in-situ recording is accretions<br />

<strong>of</strong> marine organisms, which occasionally can hide finer decoration until cleaned in the<br />

laboratory. Partial cleaning in the field might be sufficient. Also, temper studies would<br />

require that snapped samples <strong>of</strong> some or all sherds would need to be collected for analysis.<br />

317


318


CHAPTER 8:<br />

VESSEL FORM VARIABILITY AND VESSEL<br />

Introduction:<br />

FUNCTION<br />

<strong>The</strong> moderate degree <strong>of</strong> vessel brokenness <strong>of</strong> the Roviana assemblages (Chapter 5), in<br />

which many sherds are identifiable to one or more vessel parts, permits an analysis <strong>of</strong><br />

assemblage vessel form characteristics, which was carried out using a metric vessel contour<br />

approach (introduced in Chapter 1). <strong>The</strong> classification in this <strong>chapter</strong> is materialist in that<br />

it is a measurement device created by the analyst from observed variability, for the<br />

purposes <strong>of</strong> controlling to some extent for functional variation in the seriation analysis in<br />

Chapter 12.<br />

What are the sources <strong>of</strong> form variability? Are there constraints imposed by the<br />

material and method <strong>of</strong> construction? What aspects <strong>of</strong> observed form variability are related<br />

to vessel function? Is temporal variation in form constrained or directed by function, or<br />

might it fluctuate in response to some natural or social environmental condition? Which<br />

aspects <strong>of</strong> form variability might be seen as stylistic, or free to drift in a stochastic manner<br />

(Dunnell 1978, 2001)? This last question is key for the chronological aims <strong>of</strong> the analysis,<br />

while the other questions are the reverse side <strong>of</strong> the same coin. By seeking general-purpose<br />

run-<strong>of</strong> the mill vessel forms for use in seriation, some gross functional variation is removed<br />

from the mix, reducing the possibility <strong>of</strong> incorrectly seriating functional variation.<br />

Some characteristics <strong>of</strong> Roviana vessel forms are inferred from negative evidence.<br />

For example, the lack <strong>of</strong> any flat base sherds will be discussed and interpreted in this<br />

319


<strong>chapter</strong>. <strong>The</strong> main positive evidence for vessel forms is upper vessel contour variation,<br />

especially the measurements <strong>of</strong> rim angle, rim depth, rim Vcurve, neck angle, neck Vcurve,<br />

shoulder depth/Vcurve and carination angle/Vcurve (see Chapter 4 for definitions). Body<br />

sherds provided an additional line <strong>of</strong> evidence which was largely experimental, but which<br />

showed up some interesting differences between site assemblages. <strong>The</strong> precision with<br />

which these body form data are replicable has not been assessed, and there are difficulties<br />

in associating body forms with rim/neck/shoulder forms in assemblages comprising a<br />

mixture <strong>of</strong> production styles, so primary emphasis is laid on upper vessel contour for<br />

classification <strong>of</strong> vessel forms.<br />

In accordance with the theoretical discussion in Chapter 1, functional classes are<br />

kept broad, and multi-functional or general purpose forms (e.g. cooking/serving/liquid<br />

transport) are expected to dominate assemblages, on the basis <strong>of</strong> use-life theory.<br />

Subordinate (long-use-life) functional classes, for example large storage jars or ritual<br />

vessels, were also expected to be present in lower frequencies in the assemblages.<br />

Substantial form variation was expected within such broad functional classes.<br />

Sample size comes to the fore when dealing with long-use life or low-frequency<br />

functional classes, and there is no attempt here to characterize these in any detail. Analysis<br />

focuses on classification <strong>of</strong> and variability within the largest general purpose functional<br />

class. Comparison <strong>of</strong> site sample characteristics is also undertaken, using a measure <strong>of</strong><br />

sherd restriction factor, the ratio between total sherd area for a site, and rim EVE for<br />

restricted vessels. Covariation between thickness, vessel size and vessel form is examined,<br />

as it relates to the functional structure <strong>of</strong> assemblages and the characterization <strong>of</strong> vessel<br />

function. Sherd strength was not tested, but some qualitative remarks on this subject<br />

relating to assemblage functional structure are included.<br />

Methods for Inferring Vessel Form from Sherds:<br />

Rim forms alone were not regarded as a reliable guide to vessel function (this is a<br />

320


departure from the analyses <strong>of</strong> others in the general region, e.g. Summerhayes and Wickler,<br />

who in some instances regard rim forms as proxies for vessel forms (see Chapter 2). If a<br />

fundamental distinction is made between jars and bowls, it is essential that only those lip<br />

sherds which are sufficiently large to discriminate between these categories <strong>of</strong> vessel<br />

contour are taken as evidence. For very broken assemblages, neck sherds by contrast can<br />

be used to demonstrate the presence <strong>of</strong> restricted jars, as this contour is diagnostic <strong>of</strong><br />

restriction if the sherd is large enough to orient using the neck horizon. Similarly,<br />

carination sherds and sherds from the junction between flat bases and vessel sides can be<br />

used to infer carinated and flat-based vessel forms respectively. Neck/shoulder sherds can<br />

be used also to infer generally carinated form, although the form <strong>of</strong> the carination itself<br />

cannot be specified. One would expect to find carination sherds in the assemblage too in<br />

such a case. Small lip sherds cannot be inferred to represent bowls, unless independent,<br />

convincing evidence is presented to demonstrate a correlation between the particular<br />

morphology <strong>of</strong> lips and overall vessel morphology.<br />

Figure 79: Initial classification <strong>of</strong> vessel forms.<br />

321


Function and Temper:<br />

Much <strong>of</strong> the Roviana pottery is relatively strong, compared with some samples <strong>of</strong><br />

terrestrial origin (For example Reef-Santa-Cruz materials or modern Mailu pottery in<br />

storage at <strong>Auckland</strong> <strong>University</strong>). This is reflected in sherd size in some sites. A lot <strong>of</strong> the<br />

tall thick everted rims from Honiavasa for example are usually very strong, highly<br />

tempered, difficult to snap with pliers, and are quite elastic (can be dropped on the<br />

linoleum-covered concrete <strong>of</strong> the laboratory and will bounce). This seems to be evidence<br />

for either unusually strong manufacture for some production units, or preservational bias<br />

towards strong fabric (see Chapter 7). If the former were the case this would tend to<br />

support an understanding <strong>of</strong> the assemblages as predominantly cooking and/or general<br />

purpose vessels although this may relate to drying technology during pre-firing stages <strong>of</strong><br />

manufacture (drying in a torrid-zone tropical environment may require rapid fire-drying or<br />

drying in direct sunlight, necessitating increased temper concentration to reduce shrinkage-<br />

cracking (see Chapter 1). Alternately, terrestrial deposition might lead to a substantial<br />

weakening <strong>of</strong> fabric through chemical weathering, an effect not present in the sea, or<br />

perhaps immersion strengthens some sherds. Many sherds from the submerged Mulifanua<br />

site in Samoa were,<br />

“...dense and largely unaffected by age or wear. Under the petrographic<br />

microscope it was apparent that zeolites had filled in some pores. This is<br />

a phenomenon caused by sea water immersion....(Petchey 1995:57).”<br />

Some white mineral veins in some <strong>of</strong> the Roviana pottery may be a similar phenomenon.<br />

Form: Initial Nominal Classification:<br />

<strong>The</strong> following classifications <strong>of</strong> vessel form are combinations <strong>of</strong> the major nominal form<br />

attributes as described in Chapter 4. <strong>The</strong>se are preliminary classifications that are<br />

recombined into functional classes following a metric analysis <strong>of</strong> form variability within<br />

322


classes. Some are mutually exclusive, others (e.g. Form 1 and Form 6a) are<br />

complementary, in that similar vessel forms are being inferred from independent lines <strong>of</strong><br />

evidence. Some <strong>of</strong> the initial descriptive categories turn out, on further analysis <strong>of</strong><br />

variability later in the <strong>chapter</strong>, to be quite arbitrary divisions <strong>of</strong> a continuum <strong>of</strong> variation,<br />

and are lumped in a functional classification as a consequence.<br />

Form 1: Conical-to-Hyperboloid Shoulder Form and Carination:<br />

Hyperboloid form is similar to conical form but concave in vertical pr<strong>of</strong>ile (Rice 1987:219)<br />

with a corner point (Rice 1987: 218) or carination. Classification criteria in the table <strong>of</strong><br />

form data (see Chapter 4 and attached CD-ROM table “Form.db”) were “carination type”<br />

with a value “H”and “shoulder type” with a value “L”. <strong>The</strong>se forms were mainly found in<br />

the Honiavasa site (see also Figure 7, Figure 8, Figure 9 ). <strong>The</strong>y are interpreted as<br />

representing carinated restricted jars with a conical/hyperboloid shoulder, everted rims and<br />

round bases (round bases by negative evidence:- absence <strong>of</strong> flat bases in assemblages)<br />

(Figure 79). One sherd from Hoghoi had this shoulder form, and was probably broken at<br />

the carination (HG.99.061). <strong>The</strong>re was also one hard carination sherd in the Punala sample,<br />

which has not been analysed in detail. Most, though, were from Honiavasa.<br />

Form 2: Unrestricted /Weakly Restricted Rimmed Jars or Deep Bowls:<br />

This class was subdivided as follows:<br />

Form 2a:<br />

Form 2a (Figure 79, Figure 80, Figure 81) is an unrestricted jar or bowl form with a<br />

slightly excurvate rim and a thickening <strong>of</strong> the upper body giving the external appearance<br />

<strong>of</strong> a carination, without any obvious interior join, probably originating through a two-piece<br />

construction method where the excurvate rim is joined to a rounded body/base.<br />

(Classification criteria are “rim type” has value “K”, “shoulder type” does not have value<br />

“L”). This is a small group, defined mainly on their construction method and usually<br />

323


weakly restricted or unrestricted form, not easily differentiated from compound rimmed<br />

vessels. Because there were only a few <strong>of</strong> these vessels, and because decoration was<br />

confined to fingernail impression or excision <strong>of</strong> the outer rim (an attribute omitted from<br />

seriations), these sherds had no effect on the seriations.<br />

Form 2b:<br />

Form 2b (Figure 79, Figure 80) is represented by a single large quartz-calcite tempered<br />

sherd, and is thought to indicate a shallow unrestricted ellipsoidal base topped by a corner<br />

point (hard carination) and a narrow conical restriction. At the top <strong>of</strong> the conical restriction<br />

there is another corner point in the pr<strong>of</strong>ile, with a tall excurvate (hyperboloid) attached rim.<br />

Apart from the distinct corner points and the shallow base/body or lower rim, this sherd<br />

bears some resemblance to Form 2a, but was given a separate class as the exotic quartz-<br />

calcite temper in its fabric suggests a different manufacturing centre to the Form 2a sherds<br />

(suggesting the differences between this sherd and the rest <strong>of</strong> Form 2a may turn out to be<br />

more than just “noise” despite occurrence in the sample as a single sherd). Classification<br />

criteria are “rim type” with a value <strong>of</strong> “K” and shoulder type “L”.<br />

This unrestricted vessel form is sufficiently deep, rounded and well-tempered to<br />

have had similar functions to more restricted vessel forms, but equally, due to a shallow<br />

base and unrestricted access to the contents, could have fulfilled a variety <strong>of</strong> serving, non-<br />

fluid cooking, drying, washing and other functions, and would have been a fairly stable<br />

form not liable to complete capsize due to the shallowness <strong>of</strong> the base. Because it was<br />

undecorated, it was automatically excluded from all seriation analyses.<br />

Form 2c:<br />

<strong>The</strong>se were everted-rim to vertical-rim unrestricted jars or weakly restricted jars defined<br />

reasonably securely from rim/neck attributes (especially neck angle), and forming the least-<br />

restricted end <strong>of</strong> a continuum <strong>of</strong> rim/neck/shoulder variation (Figure 79, Figure 80),<br />

which included everted rim restricted jars to straight or incurvate restricted/unrestricted<br />

jars. Classification criteria were “rim type” with a value <strong>of</strong> “X”, “S” or “V” and “neck<br />

type” with a value <strong>of</strong> “T”.<br />

324


Figure 80: Examples <strong>of</strong> unrestricted Form 2 variants: Form 2a (top); Form 2b (upper<br />

middle); Form 2c (lower middle); and Form 2d (bottom).<br />

325


Figure 81: Additional Form 2a Sherds (top 6); the two decorated gambrelled vessels<br />

from Nusa Roviana are a short-rim variant <strong>of</strong> Form 2a; the lower sherd is transitional<br />

between Form 2a and Form 2d, being unrestricted.<br />

326


<strong>The</strong>se sherds do not display any morphological evidence <strong>of</strong> multi piece construction, and<br />

are inferred to be modeled from a single lump <strong>of</strong> clay by paddle and anvil techniques (or<br />

perhaps paddle and finger in some cases). Anvil impressions are common on the vessel<br />

interior below the neck, but are usually absent from the interior <strong>of</strong> the rim, presumably as<br />

a result <strong>of</strong> wiping <strong>of</strong> the rim or forming with a large anvil/tool <strong>of</strong> some sort.<br />

Form 2d:<br />

This class comprised unrestricted bowls and/or jars (Figure 79, Figure 80, Figure 81).<br />

<strong>The</strong>re were few examples, classified together by absence <strong>of</strong> any suggestion <strong>of</strong> a neck,<br />

together with excurvate rim inflecting into a globular body. <strong>The</strong>re are only two examples<br />

within this category, from two sites (Hg.10.103 and HV.5.205). Classificatory criteria in<br />

“Form.db” are: “rim type” has a value <strong>of</strong> “X” and “neck type” has a value <strong>of</strong> “N”.<br />

Form 2e:<br />

This class comprised a single sherd, suggestive <strong>of</strong> an unrestricted jar or deep bowl, with<br />

a flat thickened lip and punctate/fingernail impressed decoration (sherdA32, see images<br />

appended on CDROM). Classification criteria were “lip type” value “L” and “shoulder<br />

type” value “T”.<br />

Form 3: Inverted Rims:<br />

Form 3 vessels (Figure 79, Figure 82) were thought to be small, inverted (direct) rimmed<br />

vessels (classification as direct is tentative: there is a possibility that these are the heavily<br />

flared everted rims <strong>of</strong> small-orifice restricted vessels such as the water jar illustrated in<br />

Ross (Ross 1996:78, 2b water jar). Classification criteria were:<br />

• the lip was definitely represented on the sherd (to avoid confusion with rounded<br />

shoulders <strong>of</strong> everted-rimmed necked pots);<br />

• rim depth was sufficient to suggest absence <strong>of</strong> a neck;<br />

327


• rim form was inverted “rim type” had a value “T” in “Form.db”;<br />

• no carination was recorded (“carination type” had a value <strong>of</strong> “9"), to differentiate<br />

from one rim <strong>of</strong> this form which had a carination near the lip.<br />

Judging by the horizontal curvature <strong>of</strong> the rim, some <strong>of</strong> these sherds were from tiny<br />

vessels, about the size <strong>of</strong> an egg-cup. Vessel B.6/B.7 from Zangana was inferred to have<br />

had loop handles (Figure 83), the only example <strong>of</strong> this handle form, suggestive <strong>of</strong><br />

suspension during use, and conceivably having a specialized cooking/heating/drying<br />

function at some height above a fire which would have made support on three stones<br />

impractical. Given the uncertainties surrounding the overall vessel forms and sizes and<br />

functional implications <strong>of</strong> Form 3 these were omitted from seriations.<br />

Form 4: Shallow Carinated Bowls:<br />

<strong>The</strong>re is one example <strong>of</strong> this form, in an exotic quartz-calcite fabric ( Figure 79,Figure<br />

83).<br />

Form 5: Reworked Body Sherd:<br />

<strong>The</strong> single instance <strong>of</strong> Form 5 (Figure 79, Figure 83) was a large shallow spherical sherd<br />

with a circular lip <strong>of</strong> unusually high lip angle, with small irregularities <strong>of</strong> the lip suggesting<br />

possible reworking <strong>of</strong> a spheroidal body sherd. <strong>The</strong> sherd was found at Paniavile, and may<br />

be a potting stand or mold <strong>of</strong> some sort, or alternatively fits a description <strong>of</strong> frying pan<br />

sherds attested ethnographically in Oceania (Ross 1996:70). As this sherd was undecorated<br />

it had no effect on seriations.<br />

Form 6: Everted-rim Restricted Vessels:<br />

<strong>The</strong>se can be inferred with confidence from the numerous neck sherds indicating this form.<br />

In many cases, but not all, neck sherds can be approximately oriented by latitudinal<br />

decoration and corner point horizon to ascertain whether the vessel is clearly restricted<br />

328


Figure 82: Form 3 inverted restricted vessels.<br />

329


Figure 83: Form 3 or Form 4 (top left) and another example <strong>of</strong> a short-rim variant <strong>of</strong><br />

Form 2a (top right); the sole example <strong>of</strong> Form 3 with loop handle(s) (2 nd to top); the<br />

only confirmed Form 4 carinated bowl, in exotic temper (2 nd to bottom) and a large<br />

base sherd or frying pan, Form 5 (bottom).<br />

330


(Form 6) or unrestricted/weakly restricted (Form 2c). An analysis <strong>of</strong> metric variability<br />

within this large group <strong>of</strong> sherds was undertaken to investigate functional variation.<br />

Metric Variability Within Form 6:<br />

If orifice diameter is related to function, and fluid storage or fluid transport vessels are<br />

expected to have had small orifices, and cooking vessels larger orifices, this might produce<br />

a bimodal or multimodal distribution <strong>of</strong> orifice size (for which neck “Hcurve” measurement<br />

is used as a proxy here). Neck Hcurve for all such vessels taken together, using size<br />

intervals <strong>of</strong> 10mm, looked approximately normally distributed, but with a hint <strong>of</strong> multimodality<br />

in the tails <strong>of</strong> the distribution (Figure 84).<br />

Neck Hcurve seems largely constrained between 95mm and 155mm, although this<br />

could be a sample size or measurement effect. This makes some sense from the point <strong>of</strong><br />

view <strong>of</strong> function, since Hcurve <strong>of</strong> 95 mm is about the smallest that will admit a hand<br />

(remembering that Hcurve is measured external to the neck so neck thickness must be<br />

subtracted from Hcurve to yield orifice radius). General purpose vessels that require<br />

abrasive cleaning need to be large enough to admit a hand (this is the difference between<br />

pots or jars on the one hand and flasks or bottles on the other).<br />

To test whether this aspect <strong>of</strong> the distribution was perhaps being obscured by<br />

measurement error on small sherds, either through curve irregularity or taphonomic<br />

rounding <strong>of</strong> sherd edges, the same data were run with cases excluded where sherd area was<br />

less than 25cm2 . In this bar chart (Figure 85) sample size was much reduced, which<br />

Count<br />

30<br />

20<br />

10<br />

0<br />

100.00 200.00 300.00<br />

EXTERNAL NECK RADIUS<br />

331<br />

Bars show counts<br />

Figure 84: External neck radius, all sites, size<br />

intervals 10mm.


can be expected to introduce a greater level <strong>of</strong> jaggedness to the pattern. <strong>The</strong> pattern was<br />

clearer rather than obscured, suggesting that either measurement error was a factor on<br />

small sherds, or that the smaller vessel size class was represented principally by small<br />

sherds. Although at least one vessel style within Form 6 is represented predominantly by<br />

small sherds, as will be discussed below, this was not thought to be the explanation for the<br />

discontinuity, as neck Hcurve was generally larger than 96mm radius for that group.<br />

<strong>The</strong> relatively small number <strong>of</strong> necks measuring outside these apparent constraints<br />

were <strong>of</strong> interest. Were those necks <strong>of</strong> less than 95mm Hcurve perhaps functional variants;<br />

or evidence <strong>of</strong> multi-part construction; evidence that some potters (perhaps children) had<br />

very small hands (there is a photograph <strong>of</strong> a mother and daughter making pottery in Specht<br />

& Fields 1984 which illustrates this possibility):17); or were these merely irregularities in<br />

curvature, unrepresentative <strong>of</strong> average vessel neck radius? Were necks measuring larger<br />

than 150mm radius (the upper tail <strong>of</strong> the frequency distributions) perhaps other functional<br />

variants; unrepresentative irregularities in curvature; or a large variant specific to a<br />

particular period or production unit?<br />

Count<br />

20<br />

15<br />

10<br />

5<br />

0<br />

100.00 200.00 300.00<br />

EXTERNAL NECK RADIUS<br />

332<br />

Bars show counts<br />

Figure 85: All sites, sherds >25cm 2 , neck Hcurve intervals<br />

10mm.


Form 6a:<br />

<strong>The</strong> 95mm lower limit <strong>of</strong> external radius <strong>of</strong> the bulk <strong>of</strong> necks approaches the minimum<br />

needed for the hand <strong>of</strong> the potter to access the vessel. Only one excurvate rim sherd<br />

provides secure evidence for an external neck horizontal radius below 95mm (Figure 86),<br />

with a internal orifice radius measured as 47mm. This sherd EVE measured 25%, making<br />

irregularity <strong>of</strong> manufacture an unlikely explanation for the anomalous neck size. This sherd<br />

was from the Honiavasa site, where two-piece or multi-piece construction is evidenced by<br />

a number <strong>of</strong> slab-built carination sherds (Form 1). No neck exteriors are measurable for<br />

the Honiavasa carinated sherds, but measurement from drawn pr<strong>of</strong>iles for which the<br />

location <strong>of</strong> the CVA had been established from measurement elsewhere on the sherd<br />

suggest the following approximate data: 80mm, 90mm, 100mm and a single example <strong>of</strong><br />

less than 70mm. Together with the neck measurement <strong>of</strong> HV.04.133, these data provide<br />

two examples <strong>of</strong> orifices too small to get a large hand through, indicating that at least some<br />

vessels <strong>of</strong> this form were unlikely to have been used in any function requiring direct<br />

abrasive cleaning. <strong>The</strong>se two necks were interpreted as evidence for an everted-rim<br />

carinated vessel class with a probable use <strong>of</strong> fluid transport/storage, poorly represented in<br />

the Honiavasa assemblage, and not securely evidenced in samples from other sites<br />

(remembering there were single sherds with hard carinations at Hoghoi, Paniavile and<br />

Punala).<br />

Whether all Honiavasa vessels represented by carination sherds were so used is<br />

unknown. <strong>The</strong> effect <strong>of</strong> this vessel form on seriations is discussed in Chapter 12. Form 6a<br />

and Form 1 are regarded as possibly functionally equivalent, although the larger orifice for<br />

some Form 1 vessels could confer utility as multi purpose vessels including cooking<br />

functions.<br />

Form 6b:<br />

Of the 32 sherds which measured above 160mm external neck radius, some<br />

333


Figure 86: Form 6 “Neck Hcurve” variation; the two top sherds are too small to get a<br />

hand into (Form 6a), while the two lower sherds have head-sized or larger orifices<br />

(Form 6b).<br />

334


seemed inexplicable by curve irregularity alone but larger sherds in this category confirm<br />

the presence <strong>of</strong> a larger, less frequently discarded or preserved Form 6 vessels. Due to low<br />

numbers these do not affect the outcome <strong>of</strong> seriations in Chapter 12. Counts <strong>of</strong> vessel<br />

forms are given in Table 26.<br />

Form 6c:<br />

<strong>The</strong> general situation regarding the neck radius <strong>of</strong> everted rim restricted vessels, taking all<br />

sites together, is that the vast majority <strong>of</strong> reliable Hcurve observations fall between 95mm<br />

and 155mm radius. This range represents substantial variation in the physical size <strong>of</strong> vessel<br />

orifice, from adult hand-sized to adult head-sized, and might encompass virtually any<br />

function one might think <strong>of</strong> for a vessel <strong>of</strong> this general form, from wet/dry storage or<br />

fermentation (possibly with a tied cover) to boiling/simmering to feed a variety <strong>of</strong> group<br />

sizes, serving, drying/roasting (e.g. Canarium processing), ritual consumption or <strong>of</strong>fering,<br />

soaking, washing, transport <strong>of</strong> wet or dry materials or food processing (e.g. collection <strong>of</strong><br />

grated foodstuffs, settlement <strong>of</strong> starch grains, expression <strong>of</strong> coconut cream, processing <strong>of</strong><br />

cream into oil). <strong>The</strong> list is limited more by imagination than by vessel form. <strong>The</strong> necks tell<br />

us that these are pots rather than flasks, bowls or platters, and that these occur in a variety<br />

<strong>of</strong> neck sizes, within quite broad but strict size limits predominantly.<br />

<strong>The</strong>re is substantial form variability and decorative variability within the numerically<br />

large Form 6c (Table 26), and undoubtedly (on the evidence <strong>of</strong> rim form and evidence<br />

from body and base sherds presented below) functional variation also. <strong>The</strong>re seems little<br />

basis for separating Form 2c from Form 6c for the purposes <strong>of</strong> seriation, as assessment <strong>of</strong><br />

the degree <strong>of</strong> restriction relied on measurement <strong>of</strong> sherd orientation, which was inaccurate.<br />

Moreover the amount <strong>of</strong> restriction seemed to form a continuum, and therefore<br />

classification into Form 2c and Form 6c is arbitrary. Forms 2c and 6c combined seemed<br />

to be the best candidate <strong>of</strong> all the vessel forms on which to perform form-controlled<br />

seriations, due to large sample size, presence in all sites and decorative variability.<br />

335


Table 26: Counts <strong>of</strong> vessel form classes.<br />

Form Paniavile Hoghoi Miho Honiavasa Gharanga Nusa<br />

Roviana<br />

Not<br />

Classified<br />

336<br />

Kopo Zangana Total<br />

517 708 273 264 199 91 17 677 2746<br />

Form 1 1 1 9 1 12<br />

Form 2a 4 1 1 2 2 10<br />

Form 2b 1 1<br />

Form 2c 5 6 3 1 2 3 2 22<br />

Form 2d 1 1 2<br />

Form 2e 1 1<br />

Form 2f 1 1 2<br />

Form 3 2 1 1 6 10<br />

Form 4 1 1<br />

Form 5 1 1<br />

Form 6a 1 1<br />

Form 6b 1 1<br />

Form 6c 116 142 103 165 73 21 5 170 801<br />

Everted Rim Form Variation:<br />

Seventy-seven examples <strong>of</strong> everted rims from restricted vessels were complete in pr<strong>of</strong>ile<br />

from lip to neck, for which rim angle and rim depth were measurable. Neck angle and neck<br />

Vcurve were measurable for 60 <strong>of</strong> these sherds, and rim Vcurve for 54. <strong>The</strong>se data<br />

permitted a rough analysis <strong>of</strong> rim and neck form variability within this class (Forms 6 and<br />

2c).<br />

Rim eversion angle and rim depth for Honiavasa and Hoghoi sites illustrate the<br />

extremes <strong>of</strong> assemblage difference in rim form for the everted-rim restricted vessel class<br />

(Figure 87). Multiple models can be constructed to explain this difference: there may be<br />

two or more contemporaneous functional classes <strong>of</strong> vessel represented by groups in these<br />

data, or there may have been a gradual shift in production style over time, with overlap in<br />

occupation span accounting for Hoghoi rims being made in Honiavasa forms. <strong>The</strong>re


Figure 87: Relationship between rim angle and rim height for two sites, all<br />

measurements shown.<br />

may have been a sudden shift in trade-exchange patterning during the occupation span <strong>of</strong><br />

Hoghoi, after occupation ceased at Honiavasa, that brought a new style <strong>of</strong> pottery to the<br />

area or perhaps there was a shift in economy, encouraging the adoption <strong>of</strong> a new rim form;<br />

or alternatively a change in pottery diversity over time, where Hoghoi and Honiavasa<br />

occupations are non-contemporaneous but Honiavasa had two functional classes <strong>of</strong> pottery<br />

while Hoghoi had three.<br />

Measurement error (particularly <strong>of</strong> rim angle, given the difficulty <strong>of</strong> rim orientation<br />

measurement) is unlikely to explain the separation between these two sites, but might tend<br />

to blur the overall relationship between rim depth and rim angle in all assemblages. Using<br />

only those cases where EVE was greater than 9%, for which rim angle measurement error<br />

can be expected to be smaller, the plot <strong>of</strong> rim depth against rim angle for all sites was more<br />

expressive <strong>of</strong> a negative correlation (Figure 88).<br />

337


Figure 88: Rim angle and rim depth, all sites, filtered so that EVE is greater than<br />

9%, to reduce the effects <strong>of</strong> measurement error.<br />

This clarification <strong>of</strong> the relationship between rim depth and rim angle suggests that<br />

measurement error was creating some <strong>of</strong> the groupings in the data in Figure 87 (the<br />

greater the EVE, the more confidence we can have that the sherd is correctly oriented, on<br />

which the rim angle measurement depends). Re-examining the relationship between<br />

Honiavasa and Hoghoi rim form using EVE>9% as a filter retains a separation <strong>of</strong><br />

Honiavasa and Hoghoi rim forms, but with more linearity. It seems likely that there are<br />

technological or functional constraints on rim form, which might explain the linearity in<br />

these data. Tall rims tend to be less everted. This may be due to technical difficulty in<br />

forming a heavily-everted tall rim from a single lump <strong>of</strong> clay using paddle and anvil<br />

technique, or alternately, may reflect a continuum <strong>of</strong> intended-use variation. <strong>The</strong>se factors<br />

may combine in a temporal phyletic series, placing Honiavasa and Hoghoi at opposing ends<br />

<strong>of</strong> a temporal continuum, or may not.<br />

Outliers above this negative correlation tendency are probably slab-constructed<br />

rims, formed from an arc <strong>of</strong> clay cut from a flat slab, and the most extreme outlier <strong>of</strong> this<br />

338


sort is from the Honiavasa site. <strong>The</strong>re is other evidence for slab construction in that site,<br />

where Form 1 shoulders are most common, and where many tall everted rims are snapped<br />

from the vessel in the neck region, suggesting a weak join at the neck.<br />

<strong>The</strong> results <strong>of</strong> this analysis suggest that rim form variation <strong>of</strong> everted rim restricted<br />

jars may be constrained both by manufacturing processes and/or by functional variation,<br />

and that the effect <strong>of</strong> this variation on seriations should be assessed or controlled for.<br />

Inferring Vessel Form from Body Sherds:<br />

Curve fitting using brass template curves allowed accurate estimation <strong>of</strong> the evenness <strong>of</strong><br />

curves, and the radii <strong>of</strong> even curves. Errors <strong>of</strong> parallax that arise when using curvature<br />

charts were eliminated.<br />

<strong>The</strong> observed arrangement <strong>of</strong> interior anvil marks in latitudinal circles on larger<br />

sherds which could be oriented with confidence allowed approximate orientation <strong>of</strong> many<br />

less complete body-only sherds to be inferred. Using this method it was possible to<br />

measure the convexity <strong>of</strong> the exterior vertical pr<strong>of</strong>ile (Vcurve), and to take a second<br />

measurement normal to this curve (Hcurve). This was not necessarily a measure <strong>of</strong><br />

latitudinal curvature, and in most cases is better likened to a “great circle” curve in<br />

navigation, a curve marking the intersection <strong>of</strong> the earth’s surface and a plane passing<br />

through the earth’s centre. This curve was measured by choosing the tightest curve that<br />

would sit conformably on the sherd surface at right angles to the Vcurve measurement line.<br />

Curves which were uneven were noted but not entered as Vcurve or Hcurve.<br />

Body sherds for which anvil marks were absent or not patterned, and for which<br />

orientation could thus not be inferred, were measured in the dimension <strong>of</strong> their maximum<br />

and minimum curvatures. Where these two measurements differed they were noted in the<br />

“Notes” field <strong>of</strong> the “Form.db” table. Where they were equal (indicative <strong>of</strong> a convex<br />

spheroidal sherd surface) they were entered as Vcurve and Hcurve measurements. <strong>The</strong>re<br />

is thus a higher probability <strong>of</strong> measurement for spheroidal sherds than for cone/can forms,<br />

which is significant for the plots presented below. While all body sherds were measured<br />

in this manner, a minimum size filter <strong>of</strong> 25cm 2 was applied to the data analyzed, to reduce<br />

the possibility <strong>of</strong> sampling error due to an unknown level <strong>of</strong> curve evenness. <strong>The</strong>se<br />

339


methods are similar in overall concept to the two curvature method (Hagstrum &<br />

Hildebrand 1990). Results for all sites combined are presented in Figure 89 (Honiavasa<br />

is shown using a symbol different to the other site to show up a difference in the patterning<br />

<strong>of</strong> observations for that site.<br />

Figure 89:Body sherd curvature measurements, showing spheroidal sherds<br />

(diagonal alignment) and other canister/conical forms (e.g. s-shaped pattern<br />

<strong>of</strong> Honiavasa sherd measurements).<br />

<strong>The</strong> truncation <strong>of</strong> Vcurve data in Figure 89 at 300mm radius represents the upper limit <strong>of</strong><br />

the curve template set, and the difference between a radius <strong>of</strong> 300mm and a straight line<br />

is probably within the measurement error <strong>of</strong> the technique, given the sherd sizes under<br />

study.<br />

<strong>The</strong> other obvious patterning is concentration <strong>of</strong> the data along the line <strong>of</strong><br />

sphericity, where one curve fits the sherd in both dimensions <strong>of</strong> measurement. Most <strong>of</strong><br />

these measurements probably represent globular (spheroidal) bodied vessels <strong>of</strong> up to<br />

200mm body radius. Spheroidal sherds smaller than 100mm radius may simply be the<br />

340


smallest examples <strong>of</strong> spheroidal bodies, or may be the tightly curved bases <strong>of</strong> vessel bodies<br />

that are semi-conical in shape, approximating an upside-down parabola in external base<br />

pr<strong>of</strong>ile (see Figure 35). Sherds at the upper end <strong>of</strong> the line <strong>of</strong> sphericity are fairly gently-<br />

curving, near-flat, suggesting flattish base sherds from squat globular vessels or carinated<br />

vessels with a flattish rounded base and built-up sides.<br />

Sherds falling substantially below or above the spheroidal group are either the<br />

flatter side portions <strong>of</strong> conical lower bodies, or tightly curving (in vertical pr<strong>of</strong>ile) but<br />

horizontally moderate s<strong>of</strong>t shoulders or s<strong>of</strong>t carinations.<br />

By site, samples show some variation when plotted in this manner (Figure 90).<br />

Most sites show a cluster <strong>of</strong> spheroidal sherds above 100mm radius, extending to<br />

approximately 150-200mm. <strong>The</strong> spread in this cluster for Paniavile is probably<br />

measurement bias in the first analyzed sample, created by choosing the 150mm curve<br />

before others, and tending to accept this reading. This tendency was actively resisted by<br />

more careful curve fitting in other site assemblages. Hoghoi, Miho and Honiavasa have<br />

what are potentially pointy bases or small (less than 100mm radius) globular bodies. Miho<br />

and Zangana had outliers <strong>of</strong> low curvature.<br />

<strong>The</strong> Honiavasa sample showed a tighter clustering <strong>of</strong> spheroidal sherds than other<br />

samples, between 100 and 150mm radius, suggesting smaller-bodied globular vessels were<br />

the norm here. This sample also had a number <strong>of</strong> tight Vcurve measurements <strong>of</strong> around<br />

50mm associated with Hcurves <strong>of</strong> about 120mm suggesting a s<strong>of</strong>t carination body form,<br />

which did not occur in other sites among the body-only sherd samples (there was a s<strong>of</strong>t<br />

carination in the Hoghoi site sample).<br />

341


Figure 90: Comparison <strong>of</strong> body sherd curvature measurements by site.<br />

342


Gharanga clearly has the lowest incidence <strong>of</strong> sphericity, which is potentially significant for<br />

vessel function. A comment by the landowner that these sherds were probably indicative<br />

<strong>of</strong> pots broken during water collection from the Gharanga stream provides an additional<br />

hint that perhaps there could be some functional information in this plot. If water jars<br />

tended to be wall-sided in body shape, or biconical (unrestricted spheroid base, restricted<br />

cone or restricted hyperboloid shoulder) in shape, and the Gharanga site was a specialized<br />

water collecting site, one would expect this patterning in the data. <strong>The</strong>re are strong<br />

decorative similarities between Hoghoi and Gharanga, and yet marked dissimilarity in this<br />

plot, which may be functional difference. It should be noted that vertical-sided jars would<br />

produce restricted tall ovaloid or restricted tall elipsoid body sides with low Vcurve (Rice<br />

1987:219), but there should also be a minority <strong>of</strong> spherical sherd forms in an assemblage<br />

resulting from the breakage <strong>of</strong> bases <strong>of</strong> these vessel forms, which is the pattern seen in the<br />

Gharanga data although some <strong>of</strong> the spherical sherds that have made it through the filter<br />

are rather large at up to 200mmx200mm, indicating globular rather than pointed bases.<br />

Body Sherd Thickness: Form 7 Robust Low Curvature Body Sherds:<br />

While many sherds yielded large-radius curvature measurements, some unusually thick<br />

sherds within this category are <strong>of</strong> interest, as their large curves and robust construction<br />

suggest substantial vessels (Figure 91). Of the thicker sherds, those which yielded even<br />

curvature measurements are shown in Figure 92.<br />

343


Figure 91: Robust base sherds and cannister-shaped large body sherd (the latter having<br />

exotic quartz-calcite temper).<br />

344


<strong>The</strong>se sherds seem likely to represent large and heavy vessels. Three are cylinder-shaped<br />

in body, while the other is so flat as to be difficult to measure, with one thick tightly-<br />

Figure 92: Curvature <strong>of</strong> robust body sherds (thicker than 14mm).<br />

curved spheroidal sherd with a conical to spherical transition, likely to be the slightly<br />

pointed spherical base <strong>of</strong> a large vessel. Shott’s theory <strong>of</strong> vessel size and use-life predicts<br />

that such vessels will be under-represented in assemblages due to infrequent use/breakage<br />

(Shott 1996). In light <strong>of</strong> this well supported theory <strong>of</strong> assemblage formation, unusual<br />

sherds <strong>of</strong> these dimensions should be viewed as rare glimpses <strong>of</strong> vessel forms likely to<br />

have been a prominent feature <strong>of</strong> past economy, and <strong>of</strong> low archaeological visibility due<br />

to low breakage rates combined with low vessel sample sizes in the case <strong>of</strong> the Roviana<br />

sites.<br />

One can speculate on the functions <strong>of</strong> such vessels: storage <strong>of</strong> fluids, oils, starches<br />

or dried or otherwise preserved or fermented foods, cooking or serving for large groups<br />

on unusual occasions, or perhaps even provisioning/watering <strong>of</strong> canoes (one <strong>of</strong> these<br />

sherds was <strong>of</strong> exotic quartz-calcite temper). It seems clear that due to the paucity <strong>of</strong><br />

345


evidence for such vessels and the likelihood <strong>of</strong> functional divergence from the run-<strong>of</strong> the<br />

mill or frequently broken vessel form, that these sherds should not be used in any temporal<br />

characterization <strong>of</strong> assemblages. Absence <strong>of</strong> decoration automatically excluded them from<br />

the seriation analysis in Chapter 12.<br />

Vessel Function from Evidence for Use-Alteration:<br />

<strong>The</strong> varying degree <strong>of</strong> surface ablation <strong>of</strong> sherds from Roviana intertidal sites, evidenced<br />

by pedestalled temper grains or rounded sherd edges, suggests that surface sooting, if it<br />

was present, has generally been removed by postdepositional abrasion by water-born sand.<br />

Heat spalling was not recorded for the Roviana materials due to the difficulty <strong>of</strong><br />

distinguishing heat spalling from salt crystallisation spalling.<br />

One Form 6 vessel from Hoghoi had an encrustation <strong>of</strong> soot on the exterior (see<br />

Chapter 12), thought to have been preserved as a result <strong>of</strong> burial in s<strong>of</strong>t sediments for<br />

much <strong>of</strong> the past three millennia. One sherd from this vessel was recovered by shallow<br />

excavation, the other was recovered from the mud surface. Both had sooting. No other<br />

evidence <strong>of</strong> surface sooting was seen, which is not to say that it does not exist (soot may<br />

well be retained in surface crevices or in decorative punctations, or possibly grown over<br />

with coral). It is not assumed that sooting relates to the pot’s primary or intended use, as<br />

secondary or impromptu use over a fire without cleaning cannot be ruled out.<br />

One other sherd displayed near-surface carbon, which was dated by AMS. A plain<br />

body sherd from Honiavasa site was submitted for radiocarbon dating <strong>of</strong> a seed inclusion.<br />

<strong>The</strong> date turned out to be about 9000bp, and the laboratory also dated a thin subsurface<br />

carboniferous rind or zone near the surface <strong>of</strong> the vessel. This also yielded an age <strong>of</strong><br />

around 9000bp, suggesting that the clay incorporated organic materials <strong>of</strong> this age which<br />

had vaporised and carbonized near the surface <strong>of</strong> the vessel during firing, and that this was<br />

not use alteration.<br />

346


Twenty large body sherds were examined as a pilot study to test the incidence <strong>of</strong><br />

macro-cracking (visible to the naked eye), micro-cracking (visible under 10x<br />

magnification), pitting and spalling. Exterior and interior surfaces <strong>of</strong> sherds were compared<br />

(eleven from Miho, five from Hoghoi and four from Honiavasa). Five sherds showed<br />

interior macrocracking without exterior macrocracking. Old macrocracking <strong>of</strong> the exterior<br />

occurred in seven cases and in all but two cases this was matched by interior<br />

macrocracking. Most instances <strong>of</strong> microcracking appeared to be post-recovery as interior<br />

faces <strong>of</strong> the micro-cracks were unweathered and without marine growth, thought to be due<br />

to shrinkage on drying. Microcracking <strong>of</strong> the exterior was noted on eight sherds, matched<br />

on the interior in five cases. <strong>The</strong>re were a further three cases <strong>of</strong> interior microcracking<br />

unmatched on the exterior. Three cases <strong>of</strong> exterior spalling were noted, all from the sample<br />

<strong>of</strong> eleven Miho sherds. Three cases <strong>of</strong> interior spalling were recorded, all unmatched by<br />

exterior spalling. Two <strong>of</strong> these were from Miho and one was from Honiavasa. Exterior<br />

pitting was recorded on thirteen sherds, matched on the interior in eleven cases. Three<br />

further cases <strong>of</strong> interior pitting were recorded which were unmatched on the exterior. This<br />

pitting could have been post-depositional solution pitting <strong>of</strong> organic inclusions or<br />

something similarly s<strong>of</strong>t.<br />

<strong>The</strong>se pilot data seemed to hold little promise that any strong pattern <strong>of</strong> differential<br />

alteration <strong>of</strong> interior and exterior body sherds would be seen, and no further use-alteration<br />

data was collected. It would have been worthwhile, had the time been available, to look<br />

microscopically at indentations on body/shoulder sherds to try to find retained traces <strong>of</strong><br />

carbon. <strong>The</strong> virtual absence <strong>of</strong> evidence for use alteration should not be read as evidence<br />

for absence in the past, due to the slight polishing effect on sherd surfaces <strong>of</strong> water-borne<br />

calcite detritus in the swash zone.<br />

Secondary Smoothing <strong>of</strong> Vessel Interiors, and Interior Neck Form:<br />

347


It was noted that some interior pr<strong>of</strong>iles <strong>of</strong> restricted, everted rim vessels had a smooth<br />

upper body interior and a hard angle at the point <strong>of</strong> interior restriction (Figure 36 and<br />

Figure 93). It was assumed that this pattern was produced by secondary smoothing <strong>of</strong> the<br />

interior using a larger anvil than that used for primary forming, in contrast to those vessels<br />

having an uneven interior composed <strong>of</strong> multiple small anvil/finger impressions. Some<br />

vessels had thin walls with an even interior, but without the hard line around the interior<br />

neck. <strong>The</strong> possibilities are either that these variants represented temporal changes in<br />

potting skill/technique or that these are style or functional variation within/between<br />

assemblages. <strong>The</strong>se neck/shoulder interior variants are identified in table “Form.db” and<br />

in table “Flat.db” appended on CD-Rom.<br />

Four hundred and thirty-one sherds were recorded as having a restricted neck with<br />

both the neck and shoulder represented. Of these, 308 were <strong>of</strong> area greater than 15cm 2 ,<br />

chosen as a reasonable lower size limit, below which classification <strong>of</strong> shoulder interior<br />

evenness or neck interior form was considered insecure. Interior evenness data by site are<br />

given in Table 27. Hoghoi and Miho have the greatest abundance <strong>of</strong> even interiors, while<br />

Paniavile yielded only a small sample and Gharanga had a low abundance <strong>of</strong> even<br />

interiors.<br />

Table 27: <strong>The</strong> attribute “even/not even” by site.<br />

Site Count E Count not<br />

E<br />

348<br />

Total % E % not E<br />

Paniavile 8 14 22 36 64<br />

Hoghoi 17 19 36 47 53<br />

Miho 22 26 48 46 54<br />

Honiavasa 30 62 92 33 67<br />

Gharanga 7 28 35 20 80<br />

Zangana 23 39 62 37 63<br />

<strong>The</strong> relative abundance <strong>of</strong> hard neck interior pr<strong>of</strong>iles to not hard grouped Gharanga,


Figure 93: Hard interior neck pr<strong>of</strong>ile and even interior body/shoulder pr<strong>of</strong>ile (top); a<br />

s<strong>of</strong>ter neck interior pr<strong>of</strong>ile (middle) and uneven interior pr<strong>of</strong>ile (bottom).<br />

349


Hoghoi and Paniavile together (Paniavile had the smallest sample size, and is therefore the<br />

least reliable data), while Miho and Zangana yielded similar relative abundances. <strong>The</strong><br />

Honiavasa site was the odd one out, with intermediate relative abundance (Table 28).<br />

<strong>The</strong> relative abundance <strong>of</strong> even interiors with hard pr<strong>of</strong>iles to even interiors<br />

without hard pr<strong>of</strong>ile (see Figure 36 for definition <strong>of</strong> these attributes) grouped sites Miho,<br />

Honiavasa and Zangana together, as distinct from the Hoghoi samples. Hoghoi, and<br />

Gharanga / Kopo, if sample size is ignored, tended to have thin-walled vessels with even<br />

interiors and s<strong>of</strong>t interior pr<strong>of</strong>ile at the neck (Table 29). This tendency is highly correlated<br />

Table 28: Relative abundance <strong>of</strong> “hard neck” interior pr<strong>of</strong>iles to “not hard”.<br />

Site hard not hard total % hard % not hard<br />

Paniavile 3 19 22 14 86<br />

Hoghoi 5 31 36 14 86<br />

Miho 15 33 48 31 69<br />

Honiavasa 18 74 92 20 80<br />

Gharanga 4 31 35 11 89<br />

Nusa Rov. 5 4 9<br />

Kopo 0 4 4<br />

Zangana 18 44 62 29 71<br />

Table 29: Relative abundance <strong>of</strong> even interiors with hard neck interior pr<strong>of</strong>iles to even<br />

interiors without hard pr<strong>of</strong>ile.<br />

site even and hard even not hard total % e and s % e not s<br />

1(Paniavile)) 2 6 8<br />

2 (Hoghoi) 1 16 17 6 94<br />

3 (Miho) 9 13 22 41 59<br />

4 (Honiavasa) 14 16 30 47 53<br />

5 (Gharanga) 2 5 7<br />

6 (Nusa Rov.) 3 0 3 100<br />

7 (Kopo) 0 2 2 100<br />

8 (Zangana) 11 12 23 48 52<br />

350


with rim form (this Chapter) and decoration (Chapter 9).<br />

<strong>The</strong>se three relative abundances serve to separate Hoghoi from Miho/Zangana, and<br />

Honiavasa from all <strong>of</strong> these. Miho/Zangana have the highest relative abundance <strong>of</strong> hard<br />

neck interiors, while those site assemblages which have lower numbers <strong>of</strong> hard necks can<br />

be separated into two groups, one <strong>of</strong> which has even interiors highly correlated with hard<br />

necks, the other <strong>of</strong> which has even interiors with a s<strong>of</strong>ter neck, indicative <strong>of</strong> a different<br />

production style or technique. <strong>The</strong> low abundance <strong>of</strong> even interiors at Gharanga is <strong>of</strong><br />

interest, as decoratively Gharanga, Hoghoi and the northern end <strong>of</strong> Zangana bear many<br />

similarities (explored further in Chapters 11 and 12), and this is reflected also in Table 27;<br />

but in terms <strong>of</strong> their body sherd form these assemblages show some differences (this<br />

<strong>chapter</strong>) and this is reflected in their neck angles and rim forms also (this <strong>chapter</strong>). <strong>The</strong> low<br />

evenness count for Gharanga compared to Hoghoi may have more to do with a difference<br />

in site function than a temporal difference indicated by other aspects <strong>of</strong> vessel form,<br />

whereas the differences in evenness abundance and neck interior hard pr<strong>of</strong>ile abundance<br />

between other sites seem to correspond with stylistic differences, and are thought therefore<br />

to be more safely viewed as a temporal variant within the Roviana Lagoon region.<br />

Shoulder Form Classes <strong>of</strong> Form 6 Vessels:<br />

Two shoulder variants were recorded for restricted everted-rimmed jars: hard shoulders<br />

(see Figure 34 for a definition <strong>of</strong> these and for examples see Figure 94) and the much<br />

more common s<strong>of</strong>t shoulders (e.g. Z.11.572, Figure 94). <strong>The</strong>se variants were reminiscent<br />

<strong>of</strong> two forms recorded for the Kalinga <strong>of</strong> the Phillippines (Longacre 1991). In that case<br />

they were contemporaneous geographic variation <strong>of</strong> no functional significance within a<br />

local region. <strong>The</strong>re seems no evidence at present to suggest that these are correlates <strong>of</strong><br />

functional differences, and the Kalinga case <strong>of</strong>fers uniformitarian support for this.<br />

351


Figure 94: Hard shoulder variants <strong>of</strong> Form 6 (top and middle) as distinct from the more<br />

common s<strong>of</strong>t shoulder form (bottom).<br />

352


<strong>The</strong>se differences most likely reflect either slight differences in manufacturing technique<br />

or simply different production styles, either contemporaneous or temporal. Due to the<br />

possibility that this represents contemporaneous variation, this attribute was not used in<br />

seriations.<br />

Sherd Restriction Factor:<br />

Sherd restriction factor is a measure <strong>of</strong> the amount <strong>of</strong> surface area <strong>of</strong> sherds in an<br />

assemblage in relation to the amount <strong>of</strong> orifice represented (Smith 1985:269). Calculation<br />

<strong>of</strong> sherd restriction factor was made using EVE as a unit <strong>of</strong> rim circumference rather than<br />

cm as used by Smith. This might create some orifice-size bias in the results, but as neck<br />

Vcurve was not highly structured by site, this should not be too much <strong>of</strong> a problem. It<br />

would have been preferable to measure circumferences <strong>of</strong> lips, necks and carinations in cm<br />

as well as EVE.<br />

Table 30: Sherd restriction factor comparison <strong>of</strong> site samples.<br />

Site total sherd surface<br />

area (cm 2 )<br />

353<br />

total rim circ. (%<br />

EVE)<br />

Paniavile 9438 375 2517<br />

Hoghoi 9421 448 2103<br />

Miho 8438 274 3080<br />

Honiavasa 12487 814 1534<br />

Gharanga 4976 215 2314<br />

Zangana 11694 493 2372<br />

Sherd restriction<br />

factor (cm 2 per<br />

100%)<br />

Miho had the highest sum <strong>of</strong> sherd surface area per everted-rim vessel equivalent (Table<br />

30). Honiavasa had by far the lowest, while the other sites formed a group (in those cases<br />

where sample size permitted comparisons). <strong>The</strong> first question in relation to these data is<br />

whether collection intensity has biased the ratio <strong>of</strong> rim sherds to body sherds, and here<br />

Honiavasa has the lowest collection intensity, with Hoghoi and Zangana the highest. Miho,<br />

in spite <strong>of</strong> having an intermediate intensity <strong>of</strong> collection, not much


Table 31: Preservational bias <strong>of</strong> vessel parts as an explanation for differences in sherd<br />

restriction factor (data has the form average thickness (count) standard deviation).<br />

Site Lip<br />

Thickness<br />

(mm)<br />

Paniavile 8.21(66)<br />

2.83<br />

Hoghoi 6.32(84)<br />

1.78<br />

Miho 6.45(65)<br />

1.60<br />

Honiavasa 8.81(102)<br />

1.96<br />

Gharanga 7.11(34)<br />

1.45<br />

Zangana 7.76(80)<br />

2.37<br />

Rim<br />

Thickness<br />

(mm)<br />

7.49(141)<br />

2.12<br />

5.96(135)<br />

1.51<br />

7.21(111)<br />

1.5<br />

8.33(171)<br />

1.79<br />

7.02(49)<br />

1.56<br />

7.25(126)<br />

1.79<br />

Neck<br />

Thickness<br />

(mm)<br />

9.49(128)<br />

2.12<br />

7.72(152)<br />

2.12<br />

8.84(104)<br />

2.24<br />

10.28(173)<br />

2.34<br />

8.33(74)<br />

1.68<br />

8.93(169)<br />

2.45<br />

354<br />

Shoulder<br />

Thickness<br />

(mm)<br />

8.91(75)<br />

2.49<br />

7.14(137)<br />

2.03<br />

8.37(97)<br />

2.42<br />

9.50(146)<br />

2.48<br />

8.04(68)<br />

1.65<br />

7.84(135)<br />

2.18<br />

Body<br />

Thickness<br />

(mm)<br />

8.21(278)<br />

2.26<br />

6.67(355)<br />

1.93<br />

7.84(112)<br />

2.13<br />

8.69(139)<br />

2.06<br />

7.24(80)<br />

1.39<br />

7.03(247)<br />

1.87<br />

L:N<br />

Ratio<br />

L:B<br />

Ratio<br />

0.87 1.00<br />

0.81 0.95<br />

0.73 0.82<br />

0.86 1.01<br />

0.85 0.98<br />

0.87 1.10<br />

greater than Honiavasa, had the highest restriction factor, where for Honiavasa the figure<br />

was almost half. <strong>The</strong> intensively collected Hoghoi and Zangana sites had near average<br />

restriction factor. <strong>The</strong>re therefore seems to be almost no correlation between collection<br />

intensity and sherd restriction factor.<br />

When the ratio between average lip thickness and average body thickness was<br />

compared for site assemblages (Table 31), Miho, the site with the highest sherd restriction<br />

factor, had the lowest lip to body thickness ratio, and lip to neck thickness ratio, predicting<br />

significant postdepositional bias against lip sherds in this assemblage. Honiavasa has the<br />

thickest and most rugged lip sherds <strong>of</strong> all sites, slightly thicker on average than body<br />

sherds, <strong>of</strong>fering clear support for a taphonomic rather than vessel form / functional<br />

interpretation <strong>of</strong> restriction factor.<br />

Tall, weakly everted rim sherds may be particularly fragile and may be under-<br />

represented in assemblage EVE measurements due to failure to survive. <strong>The</strong>re is some<br />

support for such an assessment <strong>of</strong> the Miho sample and the southern area <strong>of</strong> the Zangana<br />

sample, which included some particularly tall but incomplete everted rims (Figure 95).


Chapter Summary:<br />

<strong>The</strong> primary evidence used in the identification and interptetation <strong>of</strong> vessel forms was<br />

classification <strong>of</strong> vertical contour <strong>of</strong> the upper vessel. Form variability was analysed as a<br />

means <strong>of</strong> instituting some control for functional variation in trying to identify stylistic<br />

variation, or variation subject to drift over time. As vessel form is inferred from sherds<br />

rather than observation <strong>of</strong> whole pots, epistemological issues arise in the identification <strong>of</strong><br />

form, and throughout the analysis the aim was to depend as little as possible on the form<br />

<strong>of</strong> lip/rim/sherds for information on body form.<br />

Thus analysis <strong>of</strong> vessel parts was separated into various sections, with more<br />

emphasis than is usual on body sherds, restriction factor, shoulder forms and neck/interior<br />

characteristics. Also, negative evidence was used: the absence <strong>of</strong> flat bases, stands, etc was<br />

noted. <strong>The</strong> paucity <strong>of</strong> bases identified in the Roviana data (only a few sherds were<br />

identified as bases, and these were rounded) strongly suggests that rounded, uniform<br />

thickness bases were overwhelmingly the norm. Six broad functional vessel form classes<br />

were identified through examination <strong>of</strong> upper body contour variation, while a seventh was<br />

defined on the basis <strong>of</strong> body sherd thickness and form.<br />

Forms 1-6 were defined using combinations <strong>of</strong> nominal descriptive variables. All<br />

assemblages were dominated by Form 6 (everted rim restricted vessels), identified securely<br />

by neck sherds. This vertical contour information was supplemented with metric analysis<br />

<strong>of</strong> neck and rim form, body sherd curvature and sherd restriction factor. Analysis <strong>of</strong> metric<br />

variables recording lip, rim, neck, shoulder and body curvatures, angles and thickness in<br />

some cases, resulted in re-evaluation <strong>of</strong> the nominal combinations <strong>of</strong> Forms 1-6.<br />

355


Figure 95: Tall, fragile everted excurvate rims.<br />

356


Form 6 was subdivided into 6a, 6b and 6c using neck horizontal curvature variability.<br />

Analysis <strong>of</strong> neck and orifice sizes indicated that the bulk <strong>of</strong> vessels <strong>of</strong> these forms had<br />

orifices <strong>of</strong> hand-size up to head-sized, suggesting substantial size variation. Form 6c was<br />

merged with Form 2c due to a subjective distinction between these classes, which also<br />

need not be salient to functional variation.<br />

Forms 2c and 6c had general-purpose form characteristics consistent with cooking,<br />

liquid transport, serving, storage, and processing functions. An analysis <strong>of</strong> rim variation<br />

within Form 2c/6c identified a negative correlation between rim height and eversion angle,<br />

which tied into ideas about form strength that came up in the analysis <strong>of</strong> restriction factor,<br />

and may also indicate some functional differences regarding suitability for liquid transport,<br />

stacking, and access for stirring or ladling. Short, heavily everted rims are at the<br />

stirring/ladling end <strong>of</strong> this continuum, while tall weakly everted rims tend more towards<br />

liquid transport, with the trade<strong>of</strong>f being a tendency to have a fragile thin lip (under-<br />

represented in the sample), except where slab construction is used, resulting in a more<br />

robust tall rim. Forms 2c/6c are the subject <strong>of</strong> a form/decoration analysis in the next<br />

<strong>chapter</strong>.<br />

Analysis <strong>of</strong> body sherd variability found some differences between sites. <strong>The</strong><br />

Honiavasa site was characterized as having mainly spherical bodies less than 150mm in<br />

radius, with relatively few conical or cylindrical sherds. Gharanga was at the opposite<br />

extreme, dominated by cylindrical sherd forms. Sample sizes were small for these analyses<br />

when smaller sherds were filtered out, but do illustrate the uses to which body sherds can<br />

be put in comparing samples <strong>of</strong> relatively plain potsherds.<br />

Surface ablation <strong>of</strong> sherds, while moderate in may cases, was sufficient to remove<br />

sooting in most cases, making use-alteration study difficult. Sooting was noted on one<br />

multi-purpose everted-rim Form 6c vessel (see Chapter 12). A pilot study <strong>of</strong><br />

microcracking, macrocracking and spalling found that most instances <strong>of</strong> these had<br />

probably occurred post-recovery. Some potential for soot preservation in crevices was<br />

357


noted but was not followed up due to time constraints.<br />

Form 7 comprised a small number <strong>of</strong> robust body sherds. One <strong>of</strong> these seemed<br />

likely to be the tightly rounded base <strong>of</strong> a sub-conical vessel body, while the others were<br />

wall-sided cylinder forms <strong>of</strong> relatively large horizontal radius. <strong>The</strong>se seemed likely to be<br />

rare preserved examples <strong>of</strong> high-use life large robust vessels, possibly storage vessels. One<br />

<strong>of</strong> three sherds was <strong>of</strong> exotic quartz-calcite hybrid temper.<br />

Sherd restriction factor was calculated for site samples, but was found to be subject<br />

to taphonomic bias. <strong>The</strong> part thickness data which were used to justify this conclusion<br />

draw attention to differences in lip and rim strength between sites, which created<br />

differences in the number <strong>of</strong> vessel equivalents represented by lip EVE, relative to total<br />

sherd surface-area in site assemblages. This will be shown to be significant for the analysis<br />

<strong>of</strong> decorative variation in the next <strong>chapter</strong>.<br />

A concern with sampling errors and measurement errors was maintained<br />

throughout the <strong>chapter</strong>. Various size filters were applied during analyses to systematically<br />

exclude or reduce the impact <strong>of</strong> measurement errors or poor representations <strong>of</strong> vessel<br />

characteristics.<br />

Form 1 and Forms 2c/6c are analysed decoratively in the next <strong>chapter</strong>, where the<br />

relationships between vessel form variability and decoration are investigated further. Here,<br />

the question for seriation is whether decorative variation is strongly correlated with form<br />

variation, or whether these vary independently. In the latter situation, we approach more<br />

closely a set <strong>of</strong> attributes for analysis that are not strongly structured by vessel function,<br />

but rather by style, and thus more likely to show drift with time.<br />

358


CHAPTER 9:<br />

CERAMIC DECORATIVE CLASSIFICATION AND<br />

Introduction:<br />

DECORATION/FORM VARIABILITY:<br />

Having made relatively detailed attribute descriptions <strong>of</strong> sherds, the analyst is faced with<br />

the task <strong>of</strong> refining which <strong>of</strong> the recorded variation is salient to the research questions. One<br />

crude but widely used option is to use all descriptive variability in seriation. In the worst<br />

scenario these data are used in a similarity matrix clustering approach, which will yield<br />

assemblage groupings, the significance <strong>of</strong> which may be hard to assess. <strong>The</strong> approach<br />

taken here is an exploratory analysis <strong>of</strong> covariance between variables in the data, with<br />

regard to theories <strong>of</strong> ceramic variation.<br />

Sherd decoration was recorded primarily as nominal or category data (with mark<br />

dimensions and spacing measured but mostly unused). A primary task in this <strong>chapter</strong> is to<br />

move from descriptive decorative information about sherds to information about variability<br />

<strong>of</strong> decoration at the scale <strong>of</strong> pottery vessels, as the units <strong>of</strong> behaviour across which<br />

decoration is patterned during manufacture. Some patterns are difficult to reconstruct from<br />

sherds (larger incised patterns for example). <strong>The</strong> first section deals with classification <strong>of</strong><br />

these using zone markers, which are identifiable across a range <strong>of</strong> levels <strong>of</strong> brokenness due<br />

to location at rugged corner points <strong>of</strong> the vessel pr<strong>of</strong>ile (lip, neck or carination).<br />

Simple bands <strong>of</strong> decoration, and the permutations <strong>of</strong> these across the vessel parts,<br />

provide a window into decorative structure that is, like the linear zone-markers <strong>of</strong> some<br />

<strong>of</strong> the larger patterns, insensitive to brokenness levels. <strong>The</strong> flat data structure in “Flat.db”<br />

(appended on CD) permits an evaluation <strong>of</strong> part representation and assemblage<br />

information content: the “diagnosticness” <strong>of</strong> sherds. Sherds that represent more <strong>of</strong> the<br />

vessel pr<strong>of</strong>ile tell us more about the structure <strong>of</strong> decoration across the vessel. Sample sizes<br />

359


for various part combinations are given and discussed.<br />

Covariation <strong>of</strong> decoration with vessel part is investigated using these samples, and<br />

this exploratory data analysis is used to refine ideas about the vessel production styles<br />

introduced mainly as illustrations in Chapter 3. A detailed analysis <strong>of</strong> lip<br />

impression/notching metric variability is given, allowing assessment <strong>of</strong> the temporal<br />

saliency <strong>of</strong> these decorative attributes in general, and <strong>of</strong> some widely used classificatory<br />

schemes.<br />

In a section integrating form variability with decorative variability, ideas are<br />

developed about which vessel decorative styles might be strongly constrained by or<br />

correlated with artefact form/function. Decorative types that crosscut form types can be<br />

better argued to be temporal styles (in the Dunnellian sense) rather than functional types,<br />

as there is some evidence for functional variation within the decorative class. Decorative<br />

attributes strongly correlated with particular forms may be poor attributes for seriation as<br />

there is a possibility that the analyst is seriating contemporaneous decorative variability<br />

related to function.<br />

Classification <strong>of</strong> Linear Motifs:<br />

As discussed in Chapter 1, a structural formulaic analysis is suited to simple arrangements<br />

<strong>of</strong> elements, but larger, more varied, or less structured patterns encourage the use <strong>of</strong><br />

broader classification criteria. Sample size is also an issue in making decisions about<br />

classification. <strong>The</strong> more vessels there are in a sample the more the classification can split,<br />

without running into sample size problems for counts <strong>of</strong> attributes. Conversely, where a<br />

sample is smaller, classificatory units may need to be lumped. Faced with the need to boost<br />

sample sizes <strong>of</strong> complex and highly varied linear motifs located on the rim, shoulder or<br />

body, and avoid the sparse data that would result from a splitting approach, these were<br />

categorized using two criteria<br />

• whether or not the decorative zone was delimited by a linear marker (as opposed<br />

to for example, a band <strong>of</strong> wavy deformation at the lip or a band <strong>of</strong> opposed pinch<br />

360


fingernail impression at the neck, or simply absence <strong>of</strong> a zone marker band)<br />

• whether the primary pattern was composed <strong>of</strong> single or double/triple lines)<br />

<strong>The</strong>se criteria divided the motifs into three groups, those with linear zone markers<br />

comprising double or triple lines delimiting the primary pattern (Class 1), those without<br />

linear zone markers, and with the primary pattern laid out in single lines, <strong>of</strong>ten but not<br />

necessarily with filled motifs (Class2), and unclassified motifs (Class 3).<br />

Class 1 comprised the following pattern codes: CH9, C10, C11, C12, C13, CL5,<br />

GM3, GM4, RL1. (See Table 15)<br />

Class 2 comprised the following pattern codes: CHV, CH2, CH3, CH4, CH5, CH6,<br />

CH7, CH8, C15, C17, C18, C19, C20, C23, C24, C25, CT3, CTW, GM5, GM7, PBR,<br />

RBR. (See Table 15)<br />

<strong>The</strong>se criteria were developed to express perceived differences between<br />

Honiavasa/Nusa Roviana incised/stamped motifs and all other sites.<br />

This classification-by-zone-marker owes a lot to Mead (Mead et al. 1975),<br />

although it is a lumping <strong>of</strong> Mead’s zone marker categories, and is similar to Wickler’s<br />

classification <strong>of</strong> incised motifs for the Buka Lapita-phase reef sites (Wickler 2001: 112-<br />

113), although in that case vessel corner points were viewed as zone markers, so the<br />

distinction here is slightly finer. <strong>The</strong> classification used here differs substantially from motif<br />

inventory approaches (e.g. Anson 1983, Bedford 2000), being a lumping <strong>of</strong> heterogenous<br />

motifs according to broad classificatory criteria. This has the benefit <strong>of</strong> increasing sample<br />

size for Class1 and Class 2 decoration, where a motif-based splitting approach would leave<br />

the data with almost as many classes as observations, i.e. sparse.<br />

Part Representation in the Potsherd Sample:<br />

<strong>The</strong> diagnostic value <strong>of</strong> sherds for the development <strong>of</strong> a classification utilizing decorative<br />

structure across the vessel depends on how complete a representation <strong>of</strong> vessel form and<br />

361


decoration the sherds provide. All sherds are not equal in the information they convey<br />

about decorative structure across the vessel, so sherd count alone is not an accurate<br />

representation <strong>of</strong> the sample size bearing on the analytical problem. Sample sizes for the<br />

various permutations <strong>of</strong> vessel part representation are given in Table 32, Table 33 and<br />

Table 33. <strong>The</strong>se information are given to convey the overall level <strong>of</strong> fragmentation <strong>of</strong> the<br />

sherd assemblage, particularly as it pertains to decoration and form analysis, and<br />

complement information given elsewhere on sherd area and weight, and lip EVEs.<br />

Table 32: Part representation, all sherds including necked sherds but excluding<br />

carinated and/or inverted-rim sherds.<br />

lowermost part<br />

lip rim neck shoulder body<br />

lip 22 266 91 60 14<br />

topmost rim 118 68 167 37<br />

part neck 223 210 62<br />

shoulder 233 75<br />

body 1053<br />

Table 33: Part representation for inverted-rim sherds (no neck corner point expected<br />

from morphology <strong>of</strong> sherd).<br />

lowest part<br />

rim shoulder/carination body<br />

topmost part lip 9 2 3<br />

362


Nine-hundred and fifty eight sherds were not identified by vessel part. <strong>The</strong> bulk <strong>of</strong> these<br />

would be body sherds or shoulder sherds with a convex exterior surface. For smaller sherd<br />

sizes shoulders and bodies are impossible to distinguish from each other. By site, there<br />

were differences in part representation that are thought to relate to the different breakage<br />

patterns <strong>of</strong> various pottery styles. This is illustrated by the data in Table 35.<br />

<strong>The</strong> Honiavasa site has the highest individual part counts, but the lowest relative<br />

Table 34: Carinated sherds (excluding carinated sherds with inverted rims).<br />

rim 3<br />

363<br />

lowest part<br />

carination body<br />

neck 5 2<br />

Topmost part shoulder 10 7<br />

Site Lip<br />

observation<br />

count<br />

carination 3 2<br />

Table 35: Form strength variation for different pottery styles and the effect on part<br />

representation.<br />

Rim<br />

observation<br />

count<br />

Neck<br />

observation<br />

count<br />

Shoulder<br />

observation<br />

count<br />

Paniavile 80 146 125 82 2<br />

Hoghoi 85 138 151 143 23<br />

Miho 66 115 107 98 9<br />

Honiavasa 103 171 170 148 4<br />

Gharanga 34 52 74 70 9<br />

Nusa Roviana 9 31 25 20 2<br />

Kopo 3 7 8 8 1<br />

Zangana 80 134 172 144 10<br />

total all sites 460 794 832 713 60<br />

LRNS sherd<br />

count


proportion <strong>of</strong> Lip-Rim-Neck-Shoulder sherds (LRNS). As sherd size is on average larger<br />

for this site than others, this low LRNS count is not an artefact <strong>of</strong> relative harshness <strong>of</strong><br />

taphonomic regime, but is instead interpreted as a relative weakness <strong>of</strong> the construction<br />

<strong>of</strong> necks in the site which caused many sherds to break at the neck rather than elsewhere.<br />

Similarly, the high LRNS count at Hoghoi is thought to relate to high form strength <strong>of</strong><br />

short, heavily everted rims common in this site. <strong>The</strong> low LRNS count at Miho and<br />

Paniavile have something to do with rim fragility also.<br />

Covariation Between Vessel Part and Decoration:<br />

Lips and Rims:<br />

Unbounded incision <strong>of</strong> the rim was matched by deformation into a wave form more <strong>of</strong>ten<br />

than expected from sample sizes, but this could have been sampling error as sample size<br />

for this decorative permutation was low, due to very few unbounded incision occurrences<br />

having the lip and rim in one piece (14 cases in total) (Table 36). Punctation <strong>of</strong> the rim<br />

was matched by plain lips more <strong>of</strong>ten than expected from sample sizes (15 examples where<br />

nine would have been expected from sample sizes). Plain rims showed no particular<br />

patterning with lip decoration beyond what would be expected from sample sizes.<br />

Lips and Necks:<br />

Lips and necks, being non-adjacent vessel parts, suffer from smaller sample size due to<br />

breakage (155 in total:Table 37). Plain necks do not seem to depart significantly from the<br />

incidences expected from sample sizes. This suggests lip decoration is not highly correlated<br />

with neck decoration for sherds that survive in this configuration. This does not mean that<br />

such structure was absent in the past, but suggests if it was present, it occurred<br />

364


Table 36: Lip decoration and rim decoration.<br />

Lip Decoration<br />

Class 2 punctation<br />

band<br />

365<br />

Rim Decoration<br />

unclassifie<br />

d<br />

plain Total<br />

Count<br />

unclassified 1 (7%) 3 (11%) 5 (11%) 27 (8%) 36 (8%)<br />

plain 5 (36%) 15 (54%) 16 (36%) 94 (27%) 130<br />

(30%)<br />

discontinuous deformation 2 (14%) 3 (7%) 38 (11%) 43 (10%)<br />

impression, “wav” pattern 2 (7%) 8 (2%) 10 (2%)<br />

deformation, “wav” pattern 5 (36%) 1 (4%) 5 (11%) 69 (20%) 80 (18%)<br />

fingernail pinching, single band 1 (7%) 3 (7%) 24 (7%) 28 (6%)<br />

impression, band on top edge 2 (7%) 4 (9%) 28 (9%) 34 (8%)<br />

impression, band on inner edge 3 (7%) 20 (6%) 23 (5%)<br />

impression, bands on both edges 4 (14%) 1 (2%) 8 (2%) 13 (3%)<br />

impression, bands on outer edges 1 (4%) 3 (7%) 31 (9%) 35 (8%)<br />

applied nubbins, band 1 (2%) 1 (0%)<br />

Count 14 (100%) 28<br />

(100%)<br />

on forms for which the lip and neck were easily separated.<br />

44 (100% ) 347<br />

(100%)<br />

433<br />

(100%)<br />

<strong>The</strong> low number <strong>of</strong> classified linear motifs occurring on the neck suggests this<br />

occurrence is an inaccuracy on the part <strong>of</strong> an individual potter, and that the mental<br />

template to which the potter or potters were working did not include decorating the neck<br />

in this manner (more on this below).<br />

Sherds from Honiavasa site are commonly broken at the neck. It was suspected<br />

that many <strong>of</strong> these Honiavasa necks were plain as usually there was no trace <strong>of</strong> decoration<br />

near the break. Only two lip/neck sherds represent the numerous opposed-pinch<br />

fingernail-impressed necks, suggesting that these necks are attached to a fragile rim-lip<br />

form. This is an important point for taphonomic inference and supports the idea that many


<strong>of</strong> the Honiavasa vessels were slab-constructed (see Chapter 7). <strong>The</strong> common tall rim<br />

form at Miho may have been an attempt to replicate slab-constructed tall everted rims as<br />

seen at Honiavasa, but using the one-piece method.<br />

Table 37: Lip decoration and neck decoration.<br />

Neck Decoration<br />

Lip Decoration<br />

Class<br />

2<br />

Unclassified 1<br />

(100%)<br />

punctation<br />

band unclassifie<br />

366<br />

plain band<br />

pinchin<br />

nubbins<br />

band<br />

Total<br />

2 (15%) 3 (75%) 12 (9%) 18<br />

(11%)<br />

Plain 4 (31%) 1 (25%) 45 2<br />

(34%) (100%)<br />

Deformation,<br />

discontinuous.<br />

Impressed bands,<br />

both edges, “wav”<br />

pattern<br />

Deformation<br />

“wav” pattern<br />

Pinching, single<br />

band<br />

Impression, band,<br />

top face<br />

Impression, band,<br />

inner edge<br />

Impression, band,<br />

both edges<br />

Impression, band,<br />

outer edge<br />

Column total 1<br />

(100%)<br />

1 (8%) 15<br />

(11%)<br />

2 (15%) 15<br />

(11%)<br />

1<br />

(100%)<br />

53<br />

(34%)<br />

11 (8%) 11<br />

(7%)<br />

2 (2%) 2 (1%)<br />

16<br />

(10%)<br />

10 (7%) 10<br />

(6%)<br />

17<br />

(11%)<br />

3 (23%) 6 (4%) 9 (6%)<br />

1 (8%) 8 (6%) 9 (6%)<br />

13 (100%) 4 (100%) 134 2<br />

(100%) (100%)<br />

10 (7%) 10<br />

(6%)<br />

1<br />

(100%)<br />

155<br />

(100%)


Lips and Shoulders:<br />

Here sample size is even more <strong>of</strong> a problem (71 observations in total), and a severe bias<br />

towards rugged LRNS forms is likely. No significant departures from the expected pattern<br />

based on attribute sample sizes are evident (Table 38).<br />

Table 38: Cross-tabulation <strong>of</strong> lip decoration with shoulder decoration.<br />

Lip Decoration<br />

Class<br />

2<br />

pinching,<br />

multi-band<br />

Shoulder Decoration<br />

band<br />

punctation<br />

367<br />

unclassified plain pinch<br />

band<br />

unclassified 2 1 1 1 1 6<br />

plain 1 6 2 1 19 29<br />

deformation,<br />

discontinuous<br />

deformation,<br />

“wav” pattern<br />

2 5 7<br />

6 6<br />

pinch band 3 3<br />

impressions,<br />

band, top face<br />

impressions,<br />

band, inner edge<br />

impressions,<br />

band, both edges<br />

impressions,<br />

band outer edge<br />

1 1 5 7<br />

3 1 4<br />

1 1 2 4<br />

1 4 5<br />

TOTAL 1 14 5 4 46 1 71<br />

TOTAL


Rims and Necks:<br />

Here sample size is improved to 389 due to part propinquinity (Table 39). Class 1<br />

bounded linear motifs on the rim are rare, and occur in association with single examples<br />

<strong>of</strong> unclassified neck modification, plain neck and opposed pinch fingernail-impressed band<br />

respectively. (See also sherd HV.2.297 in Figure 9 which has Class 1 decoration <strong>of</strong> the rim<br />

and a band <strong>of</strong> single fingernail impressions at the neck.) Class 2 unbounded linear motifs<br />

occur predominantly in association with fingernail-pinched necks, with a minority <strong>of</strong><br />

examples linked to plain necks also. Punctate rims are rare, and sample size is insufficient<br />

to see any patterned association with necks. Plain rims are matched more <strong>of</strong>ten than<br />

expected from sample sizes by punctate necks or by plain necks, and more rarely than<br />

expected by necks with a band <strong>of</strong> fingernail pinching.<br />

Table 39: Cross-tabulation <strong>of</strong> rim decoration and neck decoration.<br />

Neck Decoration<br />

Rim Decoration<br />

Class 2 pinching band unclas- plain pinchnub- TOTA<br />

multipunctasifiedingbins L<br />

bandtion single<br />

band<br />

Class 1 linear 1 1 1<br />

3<br />

(5%) (0%) (2%)<br />

(1%)<br />

Class 2 linear 1<br />

2 9 19<br />

31<br />

(20%)<br />

(11%) (3%) (46%)<br />

(8%)<br />

pinching,<br />

1<br />

1<br />

multi-band<br />

(100%)<br />

(1%)<br />

punctation,<br />

20 1 1 22<br />

band<br />

(7%) (2%) (33%) (6%)<br />

unclassified 1<br />

1 5 12 9<br />

28<br />

(20%)<br />

(6%) (26%) (4%) (22%)<br />

(7%)<br />

plain 3<br />

16 11 261 11 2 304<br />

(60%)<br />

(94%) (58%) (86%) (27%) (67%) (78%)<br />

TOTAL 5 (100%) 1 17 19 303 41 3 389<br />

(100%) (100%) (100%) (100%) (100%) (100%) (100%)<br />

368


Rims and Shoulders:<br />

Sherds representing both rims and shoulders numbered 243, a better sample size than lips<br />

and necks, indicating the more rugged nature <strong>of</strong> these sherds (Table 40). <strong>The</strong>re were fewer<br />

plain rims than expected that had Class 2 shoulders, fewer plain rims with shoulders having<br />

multiple bands <strong>of</strong> pinching, and fewer plain rims than expected with shoulders having a<br />

single band <strong>of</strong> pinching. Class 2 shoulders tended to have unclassified decoration, usually<br />

consisting unclassified or incomplete unbounded linear motifs (Figure 12). <strong>The</strong>re were<br />

more punctate rims with multiple bands <strong>of</strong> pinching on the shoulder than expected from<br />

sample sizes. <strong>The</strong>re were more plain rims with plain shoulders than expected from the total<br />

column percentage plain.<br />

Table 40: Cross-tabulation <strong>of</strong> rim decoration and shoulder decoration.<br />

Shoulder Decoration<br />

Rim Decoration<br />

Class 1<br />

linear<br />

Class 2<br />

linear<br />

Class 1 linear 1<br />

(10%)<br />

Class 2 linear<br />

pinching, multiband<br />

punctation,<br />

band<br />

2<br />

(20%)<br />

unclassified 4<br />

(40%)<br />

plain 1 3<br />

(100%) (30%)<br />

TOTAL 1 10<br />

(100%) (100%)<br />

Necks and Shoulders:<br />

punctati<br />

on<br />

multiband<br />

369<br />

punctat<br />

ion<br />

band<br />

unclass. Plain pinching<br />

band<br />

1 1<br />

(7%) (1%)<br />

3 16<br />

(21%) (8%)<br />

1<br />

(6%)<br />

7<br />

1 10<br />

(41%)<br />

(7%) (5%)<br />

1<br />

3 7<br />

(6%)<br />

(21%) (4%)<br />

8 5 6 156<br />

(47%) (100%) (43%) (82%)<br />

17 5 14 190<br />

(100%) (100%) (100%) (100%)<br />

3<br />

(50%)<br />

3<br />

(50%)<br />

6<br />

(100%)<br />

Total<br />

3<br />

(1%)<br />

24<br />

(10%)<br />

1<br />

(0%)<br />

18<br />

(7%)<br />

15<br />

(6%)<br />

182<br />

(74%)<br />

243<br />

(100%)<br />

Here part propinquinity benefitted sample sizes, and 449 observations pertain to this


elationship (Table 41). Of the 15 punctate necks, nearly half had multiple bands <strong>of</strong><br />

fingernail pinch on the shoulder, the rest were plain. Of the 348 plain necks, 87% were<br />

matched by plain shoulders, but there were a number <strong>of</strong> examples <strong>of</strong> multiple bands <strong>of</strong><br />

fingernail impression (14) and single bands <strong>of</strong> punctation (8) also. Plain necks occurred less<br />

<strong>of</strong>ten than expected for Class 1 and especially Class 2 linear motifs on the shoulder. A<br />

single band <strong>of</strong> pinching at the neck occurred more <strong>of</strong>ten than expected with Class 2<br />

shoulders. <strong>The</strong>se were matched in ten cases by Class 2 unbounded incision. Lateral bands<br />

<strong>of</strong> applied nubbins did not form a large sample in this analysis, but in four <strong>of</strong> seven cases<br />

were matched by Class 1 bounded incision on the shoulder, while only three cases were<br />

matched by plain shoulders.<br />

Table 41: Cross-tabulation <strong>of</strong> neck decoration and shoulder decoration.<br />

Shoulder Decoration<br />

Class Class 2 pinching punc- plain pinchin nubbin TOTAL<br />

1 linear multitationunclas- g s<br />

linear band band sified band band<br />

Class 2<br />

5 2<br />

7<br />

linear<br />

(1%) (33%) (2%)<br />

Pinching,<br />

1 1<br />

2<br />

multi-band<br />

(6%) (5%)<br />

(0%)<br />

Punctation,<br />

7<br />

8<br />

15<br />

single band<br />

(32%)<br />

(2%)<br />

(3%)<br />

Unclassified 1 3<br />

6 10 1<br />

21<br />

(12%) (18%)<br />

(24%) (3%) (17%) (5%)<br />

Plain 2 3 14 8 13 304 3 1 348<br />

(25%) (18%) (64%) (100%) (52%) (84%) (50%) (100%) (78%)<br />

Pinching, 1 10<br />

6 32<br />

49<br />

single band (12%) (59%)<br />

(24%) (9%)<br />

(11%)<br />

Band <strong>of</strong> 4<br />

3<br />

7<br />

nubbins (50%)<br />

(1%)<br />

(2%)<br />

TOTAL 8 17 22 8 25 362 6 1 449<br />

(100%) (100%) (100%) (100% (100%) (100%) (100%) (100%) (100%)<br />

Neck Decoration<br />

370


Summary <strong>of</strong> Decorative Structure Across the Vessel:<br />

Lips were more likely to be decorated than other parts (30% <strong>of</strong> lips were plain, while 78%<br />

<strong>of</strong> rims were plain, 78% <strong>of</strong> necks were plain, and 81% <strong>of</strong> shoulders were plain. Although<br />

the types <strong>of</strong> decoration present on lips was quite strongly structured by site (Table 42),<br />

there was little information to be had from the sherds themselves as to which lip decorative<br />

attributes went with which other decorative attributes, due to sherd breakage.<br />

Lip/rim sherds showed a weak relationship between Class 2 unbounded incision<br />

<strong>of</strong> the rim and wave deformation <strong>of</strong> the lip, although this could have been a sample-size<br />

error and the vast majority <strong>of</strong> wave-deformed lips which had a rim attached had plain<br />

rims. <strong>The</strong>re were no examples <strong>of</strong> unbounded incised rims with impressed decoration, but<br />

there were sample size problems with unbounded incised rims. <strong>The</strong> low number <strong>of</strong><br />

complete lip/rim sherds for this rim decorative category suggests a fragile lip/rim form.<br />

Punctate rims, by contrast, were more likely to have plain lips or impressed lips, with only<br />

one example <strong>of</strong> wave deformation. While there was one example <strong>of</strong> a punctate neck with<br />

wave-deformation <strong>of</strong> the lip, like punctate rims, punctate necks had either plain lips or<br />

impressed lips in general. In spite <strong>of</strong> small sample size, punctate shoulders showed a<br />

similar lip patterning to punctate rims and necks, and multiple bands <strong>of</strong> fingernail<br />

impression on the shoulder followed this tendency towards either plain or impressed lips<br />

also. <strong>The</strong>re was only one sherd relating unbounded incision <strong>of</strong> the shoulder to lip<br />

decoration (plain), a further indication <strong>of</strong> the fragile lip-rim form associated with<br />

unbounded incision.<br />

A band <strong>of</strong> opposed pinch fingernail impression on the lip did not occur with any but<br />

plain rims and necks, with the exception <strong>of</strong> a single sherd which had dentate-stamped<br />

bounded linear motif on the vessel body. This may relate simply to sample size, but the<br />

concentration <strong>of</strong> sherds with this lip decoration in a single site on particular vessel forms<br />

will be discussed further in Chapter 12. (This lip decoration was almost exclusively found<br />

in the Lapita-phase Honiavasa site, which is consistent with the occurrence <strong>of</strong> this<br />

decorative motif on a single dentate sherd from the Nusa Roviana site.)<br />

371


Unbounded incision <strong>of</strong> the rim was most commonly associated with a single band<br />

<strong>of</strong> opposed pinch finger-nail impression at the neck, and less commonly with a plain neck,<br />

while by contrast punctation <strong>of</strong> the rim was overwhelmingly associated with a plain neck<br />

(20 cases), with one instance <strong>of</strong> a single band <strong>of</strong> fingernail opposed pinching and one<br />

instance <strong>of</strong> a band <strong>of</strong> applied nubbins at the neck.<br />

Sherds with both rim and shoulder were not common, due to breakage, but<br />

suggested that where rims with unbounded incision were matched in the majority <strong>of</strong> cases<br />

by plain shoulders or occasionally unbounded incision on the shoulder, rims with<br />

punctation were less likely to be plain-shouldered, and where the shoulder was decorated<br />

for a punctate rim, all examples (7) were multiple bands <strong>of</strong> opposed pinch fingernail<br />

impression.<br />

Necks with a single band <strong>of</strong> opposed-pinch fingernail impression were most<br />

commonly plain shouldered, but where the shoulder was decorated in such cases, the<br />

decoration was almost always unbounded incision (10 examples). Only in three cases were<br />

plain necks associated with unbounded incision on the shoulder, and in a further three<br />

cases plain necks were associated with a single band <strong>of</strong> opposed pinch fingernail<br />

impression on the shoulder. <strong>The</strong>se latter three cases could be regarded as decorative<br />

imprecision, where the intention was to apply a band <strong>of</strong> fingernail impression to the neck,<br />

but the execution was less precise than the academic definition <strong>of</strong> body part in the present<br />

instance. Two cases where the neck had unbounded incision and the shoulder had a single<br />

band <strong>of</strong> opposed pinching were regarded as further examples <strong>of</strong> the pattern having slipped<br />

down a little.<br />

Where plain necks were associated with decorated shoulders, the decoration was<br />

most likely to be a band <strong>of</strong> punctation or multiple bands <strong>of</strong> fingernail impression. Punctate<br />

necks were sometimes matched by undecorated shoulders, and sometimes by multiple<br />

bands <strong>of</strong> fingernail impression.<br />

While sample size was small, the majority <strong>of</strong> examples <strong>of</strong> a band <strong>of</strong> applied<br />

nubbins on the neck were associated with bounded linear motifs on the shoulder, with the<br />

remainder associated with plain shoulders.<br />

372


Interpretation <strong>of</strong> Decorative Co-variation by Vessel Part:<br />

While the data presented could never prove that there are emic types in the assemblages,<br />

it is consistent with the idea that the etic decorative structure <strong>of</strong> a motif, its analytical<br />

meaning, varies according to its position on the vessel. In this view, a single band <strong>of</strong><br />

fingernail impression on the lip has a different attribute value to a band <strong>of</strong> fingernail<br />

impression at the vessel neck. Fingernail impression as multiple bands at neck or shoulder<br />

have the same value, a value which differs from either a single band <strong>of</strong> pinching at the lip,<br />

or a single band <strong>of</strong> pinching at the neck. Unbounded incision has a similar value whether<br />

it is on rim or shoulder. A single band <strong>of</strong> punctation has the same decorative value whether<br />

it occurs on rim, neck or shoulder, a case <strong>of</strong> the analyst being more specific about location<br />

on the vessel than the potter <strong>of</strong> the past. <strong>The</strong>se data supported the classification <strong>of</strong> sherd<br />

assemblages in terms <strong>of</strong> analytical decorative vessel classes, formulated in terms <strong>of</strong> the<br />

structure in the cross-tabulations given above. <strong>The</strong>se were useful for a number <strong>of</strong> purposes:<br />

to provide a univariate decorative description for spatial analysis and for easy comparison<br />

with form and fabric information below.<br />

Miho Decorative Class:<br />

Fingernail pinching <strong>of</strong> the neck, for example, commonly occurred with a plain rim or<br />

shoulder, but when decorated, almost always occurred with unbounded incision on those<br />

parts. While it was not possible with this information to classify undecorated sherds, it<br />

seemed reasonable to regard sherds with unbounded incision (neck or shoulder), and<br />

sherds with a single band <strong>of</strong> fingernail pincging at the neck, as belonging to a single group<br />

(Miho-class decoration after the Miho site where this seemed prevalent). It is possible, and<br />

likely, that the attributes which make up this grouping have independent rates <strong>of</strong> change,<br />

and a seriation using correspondence analysis in Chapter 12 uses the short list <strong>of</strong> attribute<br />

counts given in Table 42 for this reason, rather than the more normative decorative<br />

classes.<br />

<strong>The</strong> assignment <strong>of</strong> lip sherds to such a grouping is problematic due to poor<br />

373


preservation <strong>of</strong> lip-rims for this group (see Chapter 11 for a spatial approach to this<br />

problem). Wave-deformation <strong>of</strong> lips seemed to be associated with unbounded incision, but<br />

the predominance <strong>of</strong> plain rims with this lip form suggested caution, as did spatial<br />

information (Chapter 11). It is conceivable that a class <strong>of</strong> plain vessels with wave-<br />

deformation <strong>of</strong> the lip could hide within a classification that lumped all wave-deformed lips<br />

with unbounded incised rims and single-band pinched necks.<br />

Gharanga/Kopo Decorative Class:<br />

Another set <strong>of</strong> attributes that seem to group to form a style are punctation <strong>of</strong> the rim, neck<br />

or shoulder and multiple bands <strong>of</strong> fingernail pinching <strong>of</strong> the shoulder/upper body.<br />

Punctation <strong>of</strong> the rim, neck or shoulder may occur with plain adjacent parts, but where<br />

other decoration is present below this band <strong>of</strong> punctation, it was always multiple bands <strong>of</strong><br />

opposed pinch fingernail impression. This decorative association was termed<br />

Gharanga/Kopo style decoration in Chapter 3 after the sites where it is common. Rare<br />

instances <strong>of</strong> punctation <strong>of</strong> the lip occurred with a variety <strong>of</strong> decoration, and were omitted<br />

from this analysis as not highly correlated with any other decorative attribute.<br />

Attribute Frequencies:<br />

Attributes as defined in relation to vessel part above and their frequencies are given in<br />

Table 42. <strong>The</strong> data for Zangana site has been provided as a single sample, but also is split<br />

into two areas, Zangana South and Zangana North, as a result <strong>of</strong> clear large-scale spatial<br />

structure <strong>of</strong> decoration at Zangana (explained in detail in Chapter 11). <strong>The</strong> two Class 2<br />

linear motifs on the rim at Zangana North were sherds A21 and A22, found just outside<br />

the arbitrary boundary <strong>of</strong> Zangana North/Zangana South used in the preparation <strong>of</strong> Table<br />

42 (the boundary used in preparation <strong>of</strong> the attribute frequencies being in the vicinity <strong>of</strong><br />

the modern wharf at the end <strong>of</strong> the Munda-Zangana road).<br />

374


Table 42: Counts <strong>of</strong> occurrences <strong>of</strong> decorative attributes.<br />

Attributes<br />

Paniavile<br />

Hoghoi<br />

Miho<br />

375<br />

Honiavasa<br />

Sites<br />

nubbins, band at neck 2 8 1 11<br />

pinching, band at lip 5 22 1 1 28<br />

pinching, band at neck 20 4 21 1 1 2 20 1 19 69<br />

pinching, multi-band 3 10 2 11 2 1 1 28<br />

lip impression, band inner edge 2 5 1 22 3 1 1 34<br />

lip impression, band outer edge 11 7 3 4 4 2 4 2 2 35<br />

lip impression, band top face 1 10 10 6 1 6 2 4 34<br />

lip impression, band both edges 1 4 3 6 2 2 16<br />

lip impression, staggered both<br />

edges to form “wav” pattern<br />

1 5 2 2 8<br />

lip deformation, “wav” pattern 25 11 22 2 4 3 20 10 10 87<br />

lip deformation, discontinuous 8 8 12 2 1 1 1 6 2 4 39<br />

punctation, single band<br />

anywhere<br />

4 27 4 19 1 1 12 11 1 68<br />

Class 1 linear motif 16 1 17<br />

Class 2 linear motif rim 14 1 22 14 2 12 51<br />

Class 2 linear motif shoulder 10 2 3 1 8 8 24<br />

stamped decoration (dentate or<br />

wavy)<br />

6 1 7<br />

Total 104 92 100 93 56 11 2 98 34 64 555<br />

Variability <strong>of</strong> Impressed Lips:<br />

Lip notching and crenulate lips are <strong>of</strong>ten cited as a temporal markers, and in this section<br />

a sample <strong>of</strong> sherds with lip impression are analysed to examine the discreteness <strong>of</strong> these<br />

categories and to ask whether this type <strong>of</strong> simple decoration should be regarded as a<br />

homologous trait or as a simple analogous trait. In the Roviana sample, bands <strong>of</strong><br />

Gharanga<br />

Nusa Roviana<br />

Kopo<br />

Zangana combined<br />

Zangana North<br />

Zangana South<br />

Total


impression on the inner or outer and top edges <strong>of</strong> lips co-occurred with single bands <strong>of</strong><br />

punctation and multiple bands <strong>of</strong> opposed pinching lower on the vessel, and failed to occur<br />

in the single band pinched neck group. In the Honiavasa assemblage such lips were<br />

common in the absence <strong>of</strong> any punctation or multiple-band pinching, suggesting that lip<br />

impression may be a very generalized technique that may well be reinvented over time or<br />

occur across production units, and should be avoided for the purposes <strong>of</strong> seriation. <strong>The</strong><br />

presence <strong>of</strong> lip impression on late-prehistoric terrestrial plainware from Site 25, dated to<br />

1400AD provided support for this view <strong>of</strong> lip “notching” as a simple decorative analogue.<br />

Review <strong>of</strong> Lip Notching as a Temporal Marker:<br />

Specht developed a theory that Watom Lapita and Buka-Phase Lapita were characterized<br />

by a broad, deep, widely spaced notching <strong>of</strong> the lip (which looks like a v-shaped notch in<br />

Specht’s photograph), a decorative style that did not occur later, although “simple<br />

notching” was found throughout the Buka sequence (Specht 1969:218). In a similar vein,<br />

for post-Lapita pottery from New Ireland, Golson differentiated three classes <strong>of</strong> notching<br />

or crenulate decoration (v-shaped incision, parallel-sided incision/indentation, and wide<br />

stick or thumb indentations (scallops) (Golson 1991, 1992). Similarly, White and Downie<br />

differentiated several lip treatments <strong>of</strong> this sort (White & Downie 1980). Like Specht,<br />

Wickler noted rare examples <strong>of</strong> deep u-shaped impressed lip notches in site DAF (Wickler<br />

2001:120), as well as a variety <strong>of</strong> less bold notching or impression, classified in a similar<br />

way to the Roviana data above, according to position on the rim (see also Kirch et al.<br />

1991, Reeve 1989). Given that various people have classified pottery using the dimensions<br />

and form <strong>of</strong> notching or impression in various ways, it seemed that the sample <strong>of</strong> Roviana<br />

notched/impressed rims could usefully be analysed using the metric data that had been<br />

376


ecorded, to pose the question whether there were discrete classes in the Roviana data, or<br />

whether these classes are better regarded as arbitrary measurement devices, in which case<br />

they require careful metric definition. Metric data recorded for bands <strong>of</strong> lip impressions is<br />

analysed below using bivariate plots. <strong>The</strong> relationship between form and decoration is also<br />

examined, as one form attribute seems to be correlated with the location <strong>of</strong> impressions,<br />

suggesting that decorative location was an analogous decorative similarity arising from<br />

form similarity in a single attribute, rather than any general similarity in form.<br />

Lip Notching /Impression/Crenation in the Roviana sample:<br />

<strong>The</strong>re is no particular pattern to the metric data concerning banded parallel impressions on<br />

the outer lip (see Figure 96). Neither, significantly, is there any strong pattern discernable<br />

in the distribution <strong>of</strong> mark section. Not only does there seem to be little correlation<br />

between the size <strong>of</strong> impressions and the spacing <strong>of</strong> their arrangement around the rim, but<br />

there is no obvious correlation between these measurements and the shape <strong>of</strong> the marks<br />

in section, although there may be a slight tendency for narrower impressions to be more<br />

<strong>of</strong>ten v-shaped. Under Specht’s proposed Buka-Watom model, we would expect to see<br />

a positive correlation between mark width and mark spacing. This does not appear to be<br />

the case for the Roviana data, which raises the broader question for Lapita and related<br />

pottery styles whether the distinction drawn by both Specht and by Wickler is purely an<br />

arbitrary one, where occasional outliers with coarse decoration are being classified as a<br />

separate notching type, in which case early notching is simply more varied than later.<br />

Metric Data on the size and spacing <strong>of</strong> single-band impression on the inner<br />

(interior) edges <strong>of</strong> lips displays moderate positive correlation. Larger marks tend to be<br />

377


more widely spaced, while smaller, more closely spaced marks are almost always v-shaped<br />

in section (Figure 97). This latter pattern is likely to be a matter <strong>of</strong> tool choice: if the<br />

potter wished to execute a fine, closely spaced pattern <strong>of</strong> notching, then the edge <strong>of</strong> a<br />

fingernail or other fine object was used. This finer decoration was dominated by the<br />

Honiavasa site, with occasional examples from other sites too. This prevalence <strong>of</strong> the<br />

Honiavasa site in these data raises the question whether this attribute is a useful temporal<br />

marker. An argument will be put forward in the section below that it is not, that location<br />

<strong>of</strong> notching on the inner lip is related to lip orientation angle rather than directly relating<br />

to time.<br />

A similar scatterplot analysis <strong>of</strong> the size, shape and spacing <strong>of</strong> single-band<br />

impression on the top/end face <strong>of</strong> lips is reported (Figure 98). Again a moderate positive<br />

correlation can be seen between the variables, with a suggestion <strong>of</strong> a cluster <strong>of</strong> points from<br />

various sites <strong>of</strong> small, closely spaced v-section marks. This scatterplot is the closest any<br />

<strong>of</strong> the data come to showing a real distinction between fine edge-impressions and crenulate<br />

lips, but again, the overall pattern is one <strong>of</strong> a moderate positive correlation between mark<br />

width and mark spacing, with a variety <strong>of</strong> mark sections among the coarser expressions <strong>of</strong><br />

the pattern. <strong>The</strong>se is no strong patterning by site, which would be expected if the pattern<br />

was a reliable temporal marker. <strong>The</strong> virtual absence <strong>of</strong> this pattern from Gharanga and the<br />

absence <strong>of</strong> Nusa Roviana from the plot are most likely sample size effects.<br />

For sherds having double-band impression, on both edges <strong>of</strong> the lip (Figure 99)<br />

metric variables seemed weakly correlated, and mark section was either v or u, with no<br />

particular clustering in the plot. <strong>The</strong>re was no significant grouping by site.<br />

378


Figure 96: Lip impression, single band on outer edge <strong>of</strong> lip, labeled<br />

by mark section: u=u-shaped, v=v-shaped, o=oblique v, w=w-shaped,<br />

s=flat-bottomed groove.<br />

Figure 97: Lip impression, single band on inner edge <strong>of</strong> lip, labeled by<br />

mark section: u=u-shaped section, etc..<br />

379


Figure 98: Lip impression, single band on top face <strong>of</strong> lip, labeled by<br />

mark section: u=u-shaped, etc..<br />

Figure 99: Lip impression, bands on both edges <strong>of</strong> the lip. Labeled<br />

by mark section: u=u-shaped, etc..<br />

380


Conclusions, Lip Notching/Impression/Crenation:<br />

<strong>The</strong>re is a hint in the Paniavile data in Figure 96 <strong>of</strong> the pattern <strong>of</strong> coarse impression on the<br />

outer lip noted by Specht for Buka and Watom, and by Wickler for Buka, but the<br />

Honiavasa site, which is so much more evocative <strong>of</strong> Lapita proper, for various reasons<br />

discussed in Chapter 3 and Chapter 12, shows a pattern dominated by small impressions<br />

along the inner lip, with a weak correlation between mark width and mark spacing.<br />

Crenate lips, as a broad impression <strong>of</strong> the top <strong>of</strong> the lip, are not well supported as<br />

being a distinct class <strong>of</strong> decoration. Taking broad impressions as a group, without<br />

worrying too much about where on the lip they fall, there seems no pattern by site to the<br />

shape <strong>of</strong> the impressions. Either v-shaped or u-shaped can be broad and widely spaced, and<br />

crenation as a lip decoration grouping seems to have no strong structure in relation to site<br />

in the Roviana data set. <strong>The</strong>re is a continuum <strong>of</strong> variation between small impressions<br />

closely spaced and broad impressions widely spaced for impressions located on the inner<br />

lip and on the top <strong>of</strong> the lip, suggesting that a classificatory distinction between “crenate”<br />

and fine, closely spaced impression would be arbitrary rather than natural.<br />

Location <strong>of</strong> Lip Impression and Lip Form:<br />

<strong>The</strong> Roviana data is more strongly site-structured by decorative location than by size or<br />

spacing <strong>of</strong> marks. Impressions on the inner edge <strong>of</strong> the lip are common in the Honiavasa<br />

sample, and occur in some numbers in the Hoghoi sample (Table 42). Is this homologous<br />

similarity, which might place the Hoghoi punctate-neck/pinched shoulder pottery style (on<br />

which the attribute occurs at Hoghoi) closer in time to the Honiavasa site than to the Miho<br />

decorative style, or is this simple decorative technique merely an analogous similarity that<br />

would incorrectly order a seriation?<br />

381


Lip orientation angle is the sum <strong>of</strong> rim angle and lip angle, as shown in Figure 100.<br />

While lip angle was easily measured for most sherds with a well-preserved lip, rim angle<br />

was difficult to measure for all but the largest sherds, and the method used was not very<br />

precise (precision varied with EVE, but rim angle was seldom measured with a precision<br />

better than ±10 degrees. <strong>The</strong> difficulty <strong>of</strong> measuring rim angle must be borne in mind when<br />

interpreting lip orientation data.<br />

A comparison in terms <strong>of</strong> lip orientation between band <strong>of</strong> impression on the top<br />

face <strong>of</strong> the lip, band <strong>of</strong> impression on the inner edge <strong>of</strong> the lip and band <strong>of</strong> impressions on<br />

the outer edge <strong>of</strong> the lip suggests that lip orientation is influencing location <strong>of</strong> impressions<br />

(Figure 101). Although sample sizes are small and rim orientation was measured with poor<br />

precision, these data favour an explanation <strong>of</strong> the decorative similarity between Hoghoi and<br />

Honiavasa samples (bands <strong>of</strong> impression on the inner edge <strong>of</strong> the lip) as analogous rather<br />

than homologous. This interpretation is supported also by the occurrence <strong>of</strong> “bpi” pattern<br />

on Gharanga-style sherds (discussed further below).<br />

Figure 100: Calculation <strong>of</strong> lip orientation angle.<br />

382


Decorative Attributes and Vessel Form:<br />

Vessel form variability suggested subdivision <strong>of</strong> some decorative styles. Correlation<br />

between vessel form and decoration was obvious in some cases (e.g. Form 1 carinated jars<br />

and Class1 bounded linear motifs), but more difficult to approach quantitatively in others,<br />

for example in the large group <strong>of</strong> everted-rim jars. This latter group was the focus <strong>of</strong><br />

investigation <strong>of</strong> the relationship between decorative variability and form variability. <strong>The</strong><br />

classes <strong>of</strong> sherd relevant to this vessel class were everted rims, neck sherds, hard shoulders<br />

and s<strong>of</strong>t shoulders. No instances <strong>of</strong> decoration on the body <strong>of</strong> such vessels was recorded.<br />

Everted Rims:<br />

Figure 101: Lip orientation and location <strong>of</strong> impressions.<br />

In Chapter 8 a negative correlation between rim angle and rim height was suggested. Rim<br />

depth is re-plotted below for everted rims <strong>of</strong> five decorative attribute combinations<br />

(defined in detail below). Only two <strong>of</strong> the lip-rim-neck sherds with measurements<br />

pertaining to this relationship had Class 2 linear motif decoration and/or a band <strong>of</strong> pinching<br />

at the neck, a further indication <strong>of</strong> the fragility <strong>of</strong> lips associated with the Miho decorative<br />

class (Figure 102).<br />

383


Figure 102: Rim depth by decorative class.<br />

Sherds with Gharanga/Kopo-class decoration, comprising a band <strong>of</strong> punctation in the<br />

general region <strong>of</strong> the neck, and/or multiple bands <strong>of</strong> fingernail pinching on the shoulder<br />

region, can be subdivided using differences in rim form, in that pinching <strong>of</strong> the shoulder is<br />

restricted to a short-rim group, while punctation cross-cuts rim depth variability. If the<br />

short-rim variants, with or without fingernail impression, are classed as a production style<br />

on the basis <strong>of</strong> this attribute, many <strong>of</strong> the undecorated sherds would be incorporated into<br />

the class, as would the three short-rim outliers in the wave-deformed lip class in Figure<br />

102. This short-rim form, which commonly has the decorative attributes <strong>of</strong><br />

rim/neck/shoulder punctation and/or a matrix <strong>of</strong> pinching on the shoulder, was termed<br />

“Gharanga style” in Chapter 3, after the site where it was first noted.<br />

“Kopo style” is defined as those vessels with “Rimdepth” greater than 20mm, but<br />

with a band <strong>of</strong> punctation in the neck region (here style is not used in the Dunellian sense,<br />

but rather denotes a classification, as form is involved in its definition, which may have<br />

384


a functional aspect). Many vessels <strong>of</strong> this style have weakly everted rims and only slightly<br />

restricted neck, and may well be a contemporaneous functional variant <strong>of</strong> the Gharanga<br />

style. <strong>The</strong> taller, more vertical rim <strong>of</strong> the Kopo style may have been more suited to liquid<br />

transport, although access to the contents for stirring or ladling would have been less, and<br />

the form would have been less suited to use <strong>of</strong> a torque or tongs for lifting, and less suited<br />

to the use <strong>of</strong> a tied cover.<br />

Wave-deformation <strong>of</strong> the lip (fourth from left in Figure 102) only overlaps in rim<br />

depth with the Kopo style by three outliers. <strong>The</strong>se sherds (from Gharanga site) may be<br />

interpreted as homologous or phyletic links between the Gharanga/Kopo styles and wave<br />

deformation <strong>of</strong> the lip (the latter attribute weakly associated with the Miho decorative<br />

type). This is because Miho-style and Kopo-style vessels are functionally equivalent as far<br />

as can be established from form, and are likely to be temporal stylistic variants <strong>of</strong> each<br />

other. It follows that we might expect one to evolve into the other over time, and we might<br />

thus expect crossover styles forming an evolutionary continuum. <strong>The</strong> rarity <strong>of</strong> such links<br />

could result froms time-gaps in current sampling <strong>of</strong> variability.<br />

<strong>The</strong> sample <strong>of</strong> sherds with wave deformation on which the neck is still represented<br />

are typically tall, with one example <strong>of</strong> 100mmm “Rimdepth”. What cannot be seen from<br />

Figure 102 are the lip-rim sherds that are incomplete, missing the neck, and thus not<br />

measurable for rim depth.<br />

Bearing in mind the relationship between rim angle and rim depth already<br />

established in Chapter 8, the maximum measurable depths <strong>of</strong> incomplete decorated rims<br />

was <strong>of</strong> interest given the small sample <strong>of</strong> complete Miho decorative type rims, those<br />

measurable for rim angle and depth. (<strong>The</strong>re are several indications <strong>of</strong> a fragile rim form<br />

for this style, and any information pertaining to the form <strong>of</strong> what can be assumed to be<br />

many missing rims is <strong>of</strong> interest). Incomplete rims <strong>of</strong> Miho decorative type (including any<br />

385


Class 2 decorated rims that may have had wave deformation <strong>of</strong> the lip) yielded rim depths<br />

>48mm, >60mm, >65mm, >79mm, >69mm, >81mm, >76mm, >45mm, and >62mm,<br />

adding further support to the notion <strong>of</strong> a tall fragile everted rim form associated with this<br />

decorative type.<br />

Incomplete plain rims with wave deformation <strong>of</strong> the lip, which seem to have some<br />

relationship to Miho decorative type in view <strong>of</strong> five occurrences <strong>of</strong> this lip decoration on<br />

Class 2 incised lip/rim sherds, yielded the following minimum rim depth measurements: one<br />

sherd had measurable rim angle <strong>of</strong> 20 degrees, with a rim depth >65mm, and a further six<br />

tall incomplete rims yielded minimum depths <strong>of</strong> >26mm, >40mm, >47mm, >60mm,<br />

>79mm and >65mm, consistent with the data from complete rims, and consistent with the<br />

forms <strong>of</strong> Miho style decorated rims.<br />

For lips decorated with a pinched band, found mainly at Honiavasa, rim depth<br />

ranged from a minimum complete depth <strong>of</strong> 25mm (n=6) to a maximum <strong>of</strong> >65mm (n=11).<br />

<strong>The</strong>re was therefore no overlap in rim form with the Gharanga type, but overlap with the<br />

Kopo and Miho decorative classes and associated plain-rimmed wave-deformed lips.<br />

Forms <strong>of</strong> Plain Everted Rims:<br />

Sherds comprising at least the lip, rim and neck for which rim angle and rim depth were<br />

measured, and which were devoid <strong>of</strong> decoration, were plotted (Figure 103) and can be<br />

seen to be a mixture <strong>of</strong> the production styles or types discussed above. <strong>The</strong> category <strong>of</strong><br />

plain everted rim cross-cut the distinction between Gharanga and other forms, suggesting<br />

it was not a fine-grained category <strong>of</strong> stylistic or functional information, and should not be<br />

used in seriations when the aim is high temporal resolution.<br />

386


Figure 103: Rim depth and rim angle <strong>of</strong> undecorated liprim-neck<br />

sherds.<br />

Rim Form and Discontinuous Deformation <strong>of</strong> the Lip:<br />

Discontinuous deformation <strong>of</strong> the lip was present on only ten everted rim sherds that were<br />

complete from lip to neck, suggesting either a weakness at the rim/neck correlating with<br />

this decorative attribute or an unrestricted vessel form. Rim depth for this sample ranged<br />

from 21mm to 73mm (n=10), and rim angle for all measured sherds <strong>of</strong> this lip decoration<br />

ranged from 15 degrees to 50 degrees (n=10). <strong>The</strong> small samples <strong>of</strong> observations suggest<br />

that given an expanded sample there may be some overlap with the Gharanga form class,<br />

but lack <strong>of</strong> association <strong>of</strong> this lip decoration with any classified decoration on other vessel<br />

parts other than two examples <strong>of</strong> Class 2 unbounded linear motifs suggest that any such<br />

overlap was probably minor. <strong>The</strong> tendency for these sherds to be thick, chunky sherds<br />

suggests slab-construction in many cases.<br />

Everted Rim Form and Bands <strong>of</strong> Impression on the Lip:<br />

<strong>The</strong> four locational classes <strong>of</strong> lip impressed bands (bpi, bpo, bpt, bpb) were found in all<br />

sites, and analysis <strong>of</strong> decorative covariation between parts above found that these variants<br />

<strong>of</strong> lip impression were found in association with both plain sherds and with<br />

387


punctate/multi-band opposed pinch decoration on rim, neck and shoulder. <strong>The</strong> presence<br />

<strong>of</strong> these decoration classes in quantity in the Honiavasa site assemblage, from which<br />

Gharanga/Kopo decoration was absent, suggested that rather than these otherwise plain<br />

sherds being linked in some way to the Gharanga/Kopo decorative type, these were<br />

probably a set <strong>of</strong> lip attributes with a broad functional-temporal spread, possibly <strong>of</strong> little<br />

use for seriation. Sherds with this lip decoration were plotted labeled by lip decoration<br />

class (Figure 104).<br />

Although the pattern is not as marked for rim angle as it was for lip orientation<br />

angle (Figure 101), bands <strong>of</strong> impression on more everted rims tend to be on the interior<br />

edge <strong>of</strong> the lip, while on less everted rims impression tends to be on the outer edges.<br />

Impression on the top face is found across the range <strong>of</strong> rim angles, while impression on<br />

both edges ranges between 25 degrees and 55 degrees rim angle. On the rim depth scale<br />

the data fall into two groups, the sub-20mm group, which is argued above to be the<br />

Gharanga style <strong>of</strong> decoration, and the tall-rim group, in which a variety <strong>of</strong> decorative styles<br />

can be found. Lip impression, because it cross-cuts such a broad range <strong>of</strong> other decorative<br />

and form variability, is therefore regarded as likely to be analogous similarity, until proven<br />

otherwise, and is tracked carefully in the seriations in Chapter 12.<br />

Figure 104: Location <strong>of</strong> bands <strong>of</strong> lip impression in relation<br />

to rim form variability.<br />

388


Figure 105: Neck thickness comparison <strong>of</strong> Gharanga/Kopo<br />

and Miho styles/types.<br />

Neck Form and Neck Decoration:<br />

<strong>The</strong> rim-neck-shoulder area <strong>of</strong> vessels is generally robust due to form strength and part<br />

propinquinity. <strong>The</strong> Miho decorative class, most commonly identified from a single band<br />

<strong>of</strong> fingernail pinching in the neck region, was <strong>of</strong>ten seen to have a thickened neck, as in<br />

Figure 33. Gharanga-Kopo styles, on the other hand (defined as a decorative class by a<br />

band <strong>of</strong> punctation in the general region <strong>of</strong> the neck and multiple bands, or a matrix <strong>of</strong><br />

fingernail pinching in the region <strong>of</strong> the shoulder), tended to maintain a fairly constant<br />

thickness from lip to body. Metric data supporting this observation are given in Figure<br />

105.<br />

Neck thickness for Miho decorative type was on average about 2mm thicker than<br />

for Gharanga/Kopo type. Median neck Vcurve for Gharanga/Kopo decoration style was<br />

substantially smaller than for Miho decoration style (Figure 106). This information<br />

suggests that while Miho-style rims are likely to be under represented due to taphonomic<br />

389


ias, Miho-style necks are rugged as a result <strong>of</strong> form and thickness, and likely to be over-<br />

represented relative to other vessel styles/parts. This is borne out by sample sizes <strong>of</strong> the<br />

different form attributes. While only two Miho-style rims were complete from lip to neck,<br />

the count <strong>of</strong> Miho-style necks in Figure 106 is 48.<br />

Sherds which were plain on the rim, neck and shoulder, and thus not part <strong>of</strong> the<br />

Miho style by definition (with the possible exception <strong>of</strong> some wave-deformed lips; but the<br />

relationship between wave-deformation <strong>of</strong> the lip and Miho-style rims and necks is not<br />

clear) were <strong>of</strong>ten <strong>of</strong> much larger neck Vcurve than sherds <strong>of</strong> Miho style. This difference<br />

shows up in the median Vcurve <strong>of</strong> such sherds in Figure 106. Plain necks <strong>of</strong> large Vcurve<br />

were especially common in the Honiavasa site.<br />

Figure 106: Neck Vcurve for Gharanga/Kopo decoration,<br />

Miho Decoration, and plain rim-plain neck-plain shoulder<br />

sherds.<br />

390


Chapter Summary:<br />

Linear motifs were analytically classed as either those with linear zone markers or those<br />

without (with an unclassified dustbin category also). This classification is robust across a<br />

range <strong>of</strong> levels <strong>of</strong> brokenness, as the classificatory criteria are located at the edges <strong>of</strong><br />

design zones, these being the corner points <strong>of</strong> vessel pr<strong>of</strong>ile, locations which preserve well<br />

due to robust form. Patterns comprising formulaic expressions <strong>of</strong> repeated elements,<br />

together with the two linear motif classes, formed the set <strong>of</strong> decorative patterns that were<br />

used in exploratory analysis <strong>of</strong> decoration structure across vessels and in relation to vessel<br />

form variability.<br />

Class 1 motifs were largely confined to the Honiavasa assemblage, as was a lip<br />

zone marker <strong>of</strong> a band <strong>of</strong> opposed fingernail pinches. Similarly, a band <strong>of</strong> applied nubbins<br />

at the neck was virtually confined to the Honiavasa assemblage, as were sharply carinated<br />

vessel forms (Form 1), which were almost exclusively associated with Class 1 linear motifs.<br />

Deformation <strong>of</strong> the lip into a lateral wave occurred in two instances in the<br />

Honiavasa assemblage (both <strong>of</strong> these were coarse finger-impression variants <strong>of</strong> the wave<br />

form rather than lateral shear-deformation seen in most other sites). Lateral wave-<br />

deformation was more prevalent in Nusa Roviana, Paniavile and Zangana sites, with a<br />

small number <strong>of</strong> occurrences at Hoghoi and Gharanga. Wave deformation <strong>of</strong> the lip<br />

generally does not appear with punctation, with rare exceptions. While punctation was<br />

dominant at Zangana North, and Class 2 linear motifs dominated at Zangana South, wave<br />

deformation did not show this spatial patterning, occurring in several instances in both<br />

areas, suggesting possibly that wave deformation <strong>of</strong> the lip, while related to Class 2 linear<br />

motifs, may have had an extended production span relative to Class 2 linear motifs.<br />

Unlike the Honiavasa site, where the few observed instances <strong>of</strong> neck decoration are<br />

predominantly a band <strong>of</strong> applied circular nubbins (and occasionally a band <strong>of</strong> single<br />

fingernail impressions repeated), neck decoration <strong>of</strong> the other collections taken as a whole<br />

is commonly a single band <strong>of</strong> opposed-pinch fingernail impression. Where this co-occurs<br />

391


on sherds with classified decoration, the decoration is almost always unbounded linear<br />

incised (Class 2 linear). <strong>The</strong> combination <strong>of</strong> wave-deformed lip, linear incised Class 2 rim<br />

and shoulder (perhaps with some idiosyncratic applied decoration), and a band <strong>of</strong> opposed<br />

pinch decoration at the neck, is the classic expression <strong>of</strong> what I have termed the Miho style<br />

in Chapter 3 and elsewhere (Felgate 2002), but the attribute analysis in this <strong>chapter</strong><br />

suggests that a less normative approach as taken here reveals that attributes comprising<br />

this style vary independently to some extent, and may, if this variation is temporal, each<br />

have their own trajectories and rates <strong>of</strong> change.<br />

Within the Miho style there is some pattern by site regarding whether Class 2<br />

motifs occur on the rim or shoulder, with Class 2 motifs on the shoulder being almost as<br />

common as on the rim at Paniavile and Zangana South, while they occur mainly on the rim<br />

only at Miho (where Class 2 motifs occur on the shoulder it is usually motif C18 that is<br />

present). This distributional information, where the spatially distant Paniavile and Zangana<br />

South sites share a decorative similarity, while the relatively close Miho and Zangana<br />

South sites differ, suggests there may be temporal variation within the Miho style, visible<br />

in the details <strong>of</strong> attribute structure across the vessel. Accordingly, Class 2 decoration <strong>of</strong><br />

the rim and Class 2 decoration <strong>of</strong> the shoulder are treated separately in the seriations in<br />

Chapter 12.<br />

Miho-style attributes usually (but not always) occurred on excurvate rimmed<br />

restricted pots (Form 6c), and a case was made for these having a rugged neck form but<br />

a fragile rim/lip, leading to a taphonomic bias in part representation, where few complete<br />

lip/rim/neck sherds have survived.<br />

Punctation occurs predominantly as a single band in the general region <strong>of</strong> the neck<br />

(sometimes on the adjacent rim, as though the vessel was decorated while upside-down,<br />

and other times on the shoulder, clearly decorated while upright). Although a couple <strong>of</strong><br />

sherds <strong>of</strong> flared everted form had dual bands <strong>of</strong> punctation (these may have been<br />

unrestricted bowls) the single punctate band motif generally occurred on medium-height<br />

vertical-to short heavily everted rims, in the latter case <strong>of</strong>ten associated with multiple bands<br />

<strong>of</strong> pinching on<br />

392


the shoulder, forming a matrix <strong>of</strong> pinches, or a band <strong>of</strong> vertically oriented raised bars in<br />

effect, rather like some Honiavasa applied decoration. <strong>The</strong> tall punctation-only vessels<br />

were named Kopo-style, while the short-rim heavily-excurvate variants which sometimes<br />

had the matrix <strong>of</strong> pinching were called Gharanga style. Gharanga/Kopo showed a tendency<br />

for thinner necks than Miho style.<br />

<strong>The</strong>re are various applied decorations in all site assemblages, most <strong>of</strong> which seemed<br />

idiosyncratic and therefore difficult to classify. This lack <strong>of</strong> clear patterning to the applied<br />

decoration meant that in the main (with the exception <strong>of</strong> a band <strong>of</strong> circular nubbins at the<br />

neck) it was omitted from consideration in seriation analysis in Chapter 12.<br />

Homologous vs. Analogous Similarity:<br />

Analysis <strong>of</strong> lip impression/notching variability suggested there was little basis for<br />

regarding similarities between Honiavasa and Hoghoi in this regard as anything other than<br />

analogous similarity in the absence <strong>of</strong> evidence to the contrary, although the effect <strong>of</strong> these<br />

variables on seriations is considered in Chapter 12. By contrast, the Honiavasa pottery and<br />

the Miho/Gharanga-Kopo styles all share what seems likely to be a homologous trait in<br />

their neck decoration. In the Honiavasa site there were a number <strong>of</strong> instances <strong>of</strong> necks<br />

being decorated with a single band <strong>of</strong> circular nubbins. <strong>The</strong> Miho style is characterized<br />

as having a band <strong>of</strong> opposed pinching at the neck, which <strong>of</strong>ten has the appearance <strong>of</strong> a<br />

band <strong>of</strong> raised lumps <strong>of</strong> clay. In the Gharanga and Kopo styles, these are replaced by a row<br />

<strong>of</strong> punctations (almost always circular) made using a tool (access for the finger-techniques<br />

<strong>of</strong> applied nubbin and opposed pinch is restricted by the Gharanga rim form, and this<br />

could explain a change in tool used). Taken together, these three neck decorative attributes<br />

account for almost all bands <strong>of</strong> decoration at the neck or in close proximity to the neck<br />

393


(148 instances in total), and <strong>of</strong>ten occur on otherwise plain sherds, suggesting they are a<br />

primary component <strong>of</strong> a Roviana intertidal-site pottery design system, and that the three<br />

decorative techniques are three expressions <strong>of</strong> the same pattern, a row <strong>of</strong> circular elements.<br />

<strong>The</strong> three expressions <strong>of</strong> this pattern are thus likely to be homologous similarity between<br />

styles.<br />

This design feature may be phyletically related to the eyes in the Lapita faces as<br />

suggested by Spriggs (Spriggs 1990). Such an interpretation could be extended to suggest<br />

that these are either the eyes <strong>of</strong> the pot personified, or the pot as representation <strong>of</strong> person<br />

or persons, ancestral or mythical. <strong>The</strong> persistence <strong>of</strong> this design structure across the<br />

variability <strong>of</strong> the various styles suggests a temporal persistence, or, in Dunnell’s terms, a<br />

function, since the pattern is not drifting, only the technique <strong>of</strong> execution, at least at the<br />

temporal scale <strong>of</strong> the sample. <strong>The</strong> rate <strong>of</strong> change <strong>of</strong> this feature <strong>of</strong> the design system must<br />

be constrained, since it is so conservative compared to other design attributes.<br />

Eyes might have a protective ide<strong>of</strong>unction; after all, the contents, which are in most<br />

cases likely to be food, provide both sustenance and threat <strong>of</strong> poisoning, the latter either<br />

through inadvertent contamination or through malice (the distinction between these may<br />

be peculiarly modern; a result <strong>of</strong> modern microbiological knowledge). I note that previous<br />

explanations for Lapita “faces” as reviewed in the opening <strong>chapter</strong>s have tended to be<br />

soci<strong>of</strong>unctional rather than ide<strong>of</strong>unctional in that there has been a strong emphasis on ritual<br />

relating to the maintenance <strong>of</strong> Austronesian house societies and/or prestige exchange. Here<br />

the suggestion is put forward that for the Roviana pottery at least, a more prosaic efficacy<br />

may be being attempted, simply protecting the contents through ritualistic manufacture<br />

including the application <strong>of</strong> decorative eyes. Where the house-ancestor and prestige<br />

exchange theories suggest male manufacturers, an ide<strong>of</strong>unctional explanation <strong>of</strong> the sort<br />

suggested here is more consistent with utilitarian production by women.<br />

In this <strong>chapter</strong> a spatial division <strong>of</strong> style between Zangana North and Zangana<br />

South has been introduced, which is examined in more detail in Chapter 11, prior to the<br />

ceramic seriations in Chapter 12.<br />

394


Introduction:<br />

CHAPTER 10:<br />

LITHICS<br />

<strong>The</strong> main focus <strong>of</strong> this research has been ceramics, but the intertidal/shallow water lag<br />

deposits <strong>of</strong> resistant materials have the potential to provide good samples <strong>of</strong> lithic artefacts<br />

and manuports. <strong>The</strong> analysis <strong>of</strong> lithics had several practical applications in meeting the<br />

overall methodological objectives <strong>of</strong> this research. Firstly, the form <strong>of</strong> lithic artefacts has<br />

the potential to slot the Roviana materials into an existing Culture-Historical framework,<br />

providing traditional and useful independent evidence for chronological corroboration,<br />

however coarse.<br />

Secondly, this study has a central aim <strong>of</strong> providing a methodological scoping<br />

exercise for these intertidal sites: how might we proceed in seeking to write prehistory<br />

from the materials at hand? <strong>The</strong> intertidal artefact/manuport scatters have a conspicuous<br />

lithic component, due to the rugged nature <strong>of</strong> lithics. We have so little preserved to work<br />

with that it is impossible to ignore the lithic component <strong>of</strong> assemblages in seeking to<br />

develop applied archaeological method. <strong>The</strong> preserved lithic material can be used in<br />

sourcing studies, in behavioural studies, and in taphonomic studies, and this <strong>chapter</strong><br />

provides pilot studies along each <strong>of</strong> these lines <strong>of</strong> enquiry. Substantive results are not earth<br />

shattering, but the attempt is made in order to think about what sort <strong>of</strong> information might<br />

usefully be acquired from these sites and how it might be used.<br />

Although Reeve reports an obsidian point from Paniavile, and Wickler found<br />

copious quantities <strong>of</strong> obsidian on the Lapita reef sites <strong>of</strong> Buka, no obsidian, and only two<br />

small flakes <strong>of</strong> chert were found in the present study. This may be an observer bias, or a<br />

different setting which alters the visibility <strong>of</strong> these materials, but the expectation that these<br />

395


would be present was not met during extensive searching with the aim <strong>of</strong> locating obsidian.<br />

Shell adzes have been numerous in some early ceramic sites (more than 200 from<br />

AN-6 on Anuta) (Kirch & Rosendahl 1973:67). While there are many small Tridacna<br />

adzes littering the surface <strong>of</strong> some late-prehistoric sites <strong>of</strong> the Roviana area, none have<br />

been found so far from the intertidal sites, other than the three Tridacna adzes reportedly<br />

recovered from Paniavile by villagers, illustrated by Reeve (Reeve 1989). <strong>The</strong> lack <strong>of</strong> these<br />

in all recent collections may be either a preservation, observer, or visibility bias or a<br />

behavioural difference between the occupants <strong>of</strong> the collection sites in the past (we cannot<br />

distinguish between these possibilities on current evidence). In order to compare like with<br />

like, in terms <strong>of</strong> raw material properties and manufacturing constraints, no comparisons are<br />

made between the stone adzes recovered from the Roviana intertidal sites and the shell<br />

adzes recovered from early ceramic sites elsewhere.<br />

A section about water-rounded and fractured lithic manuports attempts a<br />

petrographic classification based on a macroscopic type series. Petrographic results 1 allow<br />

reclassification into broad source groupings. <strong>The</strong>se groupings will be used in Chapter 11<br />

to investigate whether manuport source classes are spatially structured at Hoghoi.<br />

Additionally, manuport size frequency distribution is used to make some behavioural<br />

inferences about procurement strategies and uses <strong>of</strong> these. <strong>The</strong> results are valuable in that<br />

they raise a series <strong>of</strong> questions on these topics which show the value in focusing this sort<br />

<strong>of</strong> attention on these materials. Also, many <strong>of</strong> the nondescript fragments that looked as<br />

though they held little archaeological value individually turned out to be more exciting<br />

when analyzed petrographically, and suggest that intensive analysis <strong>of</strong> all lithic materials<br />

can pay major dividends for this type <strong>of</strong> site. A wealth <strong>of</strong> information content is there,<br />

1 Petrographic descriptions were transcribed by the author from verbal descriptions<br />

generously provided by Dr Robin Parker, <strong>of</strong> the Geology Department, <strong>University</strong> <strong>of</strong><br />

<strong>Auckland</strong>.<br />

396


awaiting detailed analysis.<br />

Formation processes are a major focus <strong>of</strong> the ceramic analysis. <strong>The</strong> lithics, like the<br />

ceramics, can provide information on formation processes through brokenness,<br />

completeness and remaining use life. For lithic artefacts, brokenness and completeness can<br />

be used in the same way as ceramics, to estimate a breakage population, except that<br />

artefacts like adzes have bilateral symmetry rather than circular symmetry about a central<br />

axis. <strong>The</strong> difference between a breakage population and the extant deposit may inform on<br />

formation processes if a sufficient sample is available to allow an estimate <strong>of</strong> breakage<br />

population within useful limits. This was not the case with the sample <strong>of</strong> adzes obtained,<br />

but an expanded sample would potentially allow this sort <strong>of</strong> approach. Remaining use life<br />

can be assessed for those adzes that are recovered whole, by dimensional comparison with<br />

unused preforms. Where the degree <strong>of</strong> resharpening is low, a substantial use life remains,<br />

and if this is the dominant pattern, site abandonment or destruction by a sudden natural<br />

disaster or attack is a possibility. Where adzes are all re-sharpened to stubby dimensions<br />

with little remaining use life other than recycling to smaller tools, gradual discard through<br />

unexceptional processes is more likely. Some rudimentary comments can be made on the<br />

Roviana materials in this regard, as one unused preform from a local landowner/collector<br />

was recorded.<br />

Sourcing studies can only be taken so far within the limited resources <strong>of</strong> a thesis,<br />

but petrographic descriptions given in this <strong>chapter</strong> illuminate a diversity <strong>of</strong> geological raw<br />

materials. <strong>The</strong> geological origins from which these were sourced and the uses to which<br />

these were put are limited by little more than the imagination at present, and it is hoped<br />

that these data will demonstrate the desirability <strong>of</strong> making these materials the object <strong>of</strong> a<br />

sustained program <strong>of</strong> research and cultural heritage resource management.<br />

397


Review:<br />

Formation theory suggests that unlike utilitarian pottery, stone axes may have long use-<br />

lives and low discard rates (Shott & Sillitoe 2001:276), leading to an expectation that<br />

sample size will be smaller for stone axes than for pots. It is assumed here that the same<br />

holds for stone adzes. Turner’s discussion <strong>of</strong> adze use-life (Turner 2000: 231-296) does<br />

not cover the relative representation <strong>of</strong> adzes and other materials in archaeological<br />

samples, but rather focuses on the various use-life states <strong>of</strong> recovered adzes and<br />

experimental adzes. <strong>The</strong> conclusion that, “Both experimental and archaeological data<br />

support the probability that adzes had long use lives and underwent considerable<br />

morphological changes as a result <strong>of</strong> intensive curation (Turner 2000:296).” supports that<br />

assumption.<br />

In line with this expectation, few stone adzes have been recovered from Lapita-age<br />

sites in Near Oceania. Four are reported from site SAC at Watom (Green & Anson 2000b:<br />

59), and “...a few flakes... (Golson 1991:255) are reported from Lasigi. <strong>The</strong>re is no<br />

mention <strong>of</strong> stone adzes in any <strong>of</strong> the Arawe Lapita site reports (Gosden 1989, 1991a, b,<br />

Gosden et al. 1989, Gosden & Webb 1994). Few are reported from excavations in the<br />

remote Oceanic islands <strong>of</strong> the southeast Solomons or Vanuatu, with the exception <strong>of</strong> the<br />

Reef/Santa Cruz sample (much <strong>of</strong> which may have been surface collected, and excavated<br />

area was large). On Anuta, for example, no stone adzes were found either from<br />

excavations at AN-6 or from surface collections, while more than two hundred shell adzes<br />

were recovered (Kirch & Rosendahl 1973:66). Thirteen stone adzes or flakes from adzes<br />

were reported as analysed petrographically from the Lapita sites <strong>of</strong> the Reef/Santa Cruz<br />

Islands (Green 1976a:259), although it is not known whether these are all from separate<br />

adzes; the number <strong>of</strong> adzes represented by that sample may be less than thirteen. <strong>The</strong>re are<br />

also a number <strong>of</strong> other as-yet unreported stone flakes (most with signs <strong>of</strong> grinding) in the<br />

398


Reef-Santa-Cruz collections, some <strong>of</strong> which are clearly fragments <strong>of</strong> adzes (Sheppard<br />

2003: pers. comm.). Apparently no stone adzes were recovered during the extensive<br />

program <strong>of</strong> survey and excavation <strong>of</strong> Lapita and post-Lapita sites on Mussau, as there<br />

seems to be no reporting <strong>of</strong> any such items in the most recent volume on the subject,<br />

although no conspicuous mention is made <strong>of</strong> this absence in the report (Kirch 2001).<br />

In Vanuatu, the best Lapita samples at present are from surface sites on Malo.<br />

Hedrick found two stone adzes here (one plano-lateral and one plano-convex) (Hedrick<br />

n.d). Ten shell adzes were recovered from the Atanoasao site on Malo, seven from surface<br />

collection and three from excavation, but no stone adzes were found (Galipaud 1998).<br />

Recent intensive testpitting on a number <strong>of</strong> early Vanuatu sites (Bedford 2000: 205-206)<br />

recovered one Tridacna adze from Bedford’s early period (3000-2800bp), from the<br />

Arapus site, while a number <strong>of</strong> shell adzes have been recovered from later contexts,<br />

together with a single rectangular-sectioned stone adze from Bedford’s 2800-2500bp<br />

period.<br />

In contrast to this dearth <strong>of</strong> evidence from excavated Lapita sites on land in island<br />

Melanesia, there has been a relative flood <strong>of</strong> stone adzes and flakes from reef sites in Near<br />

Oceania, which may say more about the nature <strong>of</strong> the sampling process than any behavioral<br />

differences (large areas are being sampled by the surface collections, as done also by Green<br />

in the extensive surface collections and area-excavations in the Reef-Santa Cruz sites).<br />

Three complete and four fragmentary stone adzes were recovered from Paniavile by local<br />

collectors (Reeve 1989:55), which were either quadrangular or Plano-convex in cross<br />

section.<br />

Green and Davidson’s Samoan Adze typology (Green & Davidson 1969) as<br />

recently modified by Wickler for a relatively large Island Melanesian sample (Wickler<br />

2001:181) provides the classificatory units for Wickler’s study <strong>of</strong> Buka stone adzes. Green<br />

and Davidson based their Samoan typology on previous schemes <strong>of</strong> Buck’s and Poulsen’s,<br />

399


ut with more emphasis on cross section attributes, rather than the emphasis on attributes<br />

<strong>of</strong> the butt, lugs, or curvature <strong>of</strong> the back as in Duff’s scheme (Green & Davidson 1969).<br />

Green and Davidson’s type IVa and type V were characteristic <strong>of</strong> early Samoan ceramic<br />

sites, and were rare in surface collections. <strong>The</strong>se are similar in cross section, being flat on<br />

one surface and convex-to-trapezoidal on the other, but type IVa is sharpened with the<br />

edge at the flat side, producing a flat cut, while type V is sharpened with a horseho<strong>of</strong> -<br />

shaped edge along the convex side, which produces a gouge-shaped cut.<br />

Green and Davidson’s types VI, VII and VIII are all triangular-sectioned,<br />

sharpened either way to produce either a narrow straight edge or a narrow gouge, and<br />

some <strong>of</strong> these are large and heavy, but many are small and fragmentary. Type VI was<br />

thought to be present throughout the Samoan sequence. Type VII was a narrow deeper<br />

version <strong>of</strong> type V, with a prominent narrow gouge-shaped cutting edge, as though for<br />

producing grooves or rebates, and was found only in early contexts. Type VIII is rare in<br />

the Samoan sample, and although only known from surface contexts, was inferred to be<br />

an early form from its rarity. This was a heterogeneous type, most subtypes <strong>of</strong> which had<br />

a straight cutting edge and a triangular -sectioned back, and were a broad triangle in cross-<br />

section (broader than type VI).<br />

Wickler looked at 18 adzes and 19 triangular-sectioned flaked stone tools from the<br />

Buka reef sites. Wickler departed from Green and Davidson’s typology in that fragments<br />

<strong>of</strong> triangular-sectioned flaked stone tools were classified separate to the adzes (Wickler<br />

2001:175, 177), when these may well have been broken fragments <strong>of</strong> small type VI adzes.<br />

Wickler describes ten <strong>of</strong> these as flakes <strong>of</strong> dark green lithified mudstone, and nine as<br />

andesite, both materials used in the manufacture <strong>of</strong> the adzes in that sample. Five <strong>of</strong> these<br />

had ground or polished margins. Both factors suggest that at least some <strong>of</strong> these fragments<br />

could be classified as small Green and Davidson type VI adze fragments. Whether the term<br />

adze is appropriate to these small narrow forms or whether they were used as chisels<br />

400


is unimportant in the current context, where we are dealing with small fragments only, and<br />

are limited to discussion largely <strong>of</strong> differences in cross-section.<br />

Roviana Lithic Artefacts:<br />

Lithic artefacts from the intertidal included seven relatively complete adzes (Figure 107,<br />

Figure 108, Figure 109, Figure 110), a number <strong>of</strong> small fragments <strong>of</strong> what were probably<br />

adzes (not illustrated), two chert flakes, a worked sandstone slab, notched as though for<br />

an anchor and a number <strong>of</strong> small cobbles with depressions on the edges and ends, semi-<br />

cubic in form, that are interpreted as Canarium hammers (Figure 111). <strong>The</strong>se differ from<br />

an ethnographically known hafted nut hammer found widely on land in the region, and are<br />

identified by the use-indents on the six faces <strong>of</strong> the cube. <strong>The</strong>y are typically composed <strong>of</strong><br />

a relatively s<strong>of</strong>t rock that is inferred to have indented easily with use, compared to the<br />

hafted variety, which is not to say these were too s<strong>of</strong>t to be functional.<br />

Petrographic descriptions <strong>of</strong> lithic artefacts are given in Table 43. <strong>The</strong>se data<br />

suggest a diversity <strong>of</strong> lithic resources were exploited for the manufacture <strong>of</strong> artefacts,<br />

including at least two sources <strong>of</strong> adze rock, one represented by A.12, A.31 and A.6, the<br />

other, more diverse group represented by the remainder <strong>of</strong> the adze fragments. <strong>The</strong> second<br />

petrographic grouping includes substantial differences in the surface appearance <strong>of</strong> the<br />

stone, with one adze and one fragment from Hoghoi being glossy black in appearance,<br />

while other fragments from Zangana and Miho within this broad petrographic grouping<br />

were more weathered in appearance, in a variety <strong>of</strong> hues. <strong>The</strong>se surface differences suggest<br />

that if further analysis was carried out to characterize diagenesis within this grouping it<br />

could be further subdivided on a systematic basis.<br />

401


Figure 107: Planilateral sectioned adze from Hoghoi initial surface collection (top),<br />

plano-convex-sectioned type V adze from Zangana (middle), and planilateral adze<br />

fragment from Zangana (bottom).<br />

402


Figure 108: Images <strong>of</strong> the adzes shown in the preceding illustration.<br />

403


Figure 109: Green “type VI” or “type VIII” triangular-section adze fragment from<br />

Miho (top); butt-end <strong>of</strong> a plano-lateral-sectioned adze from Zangana South (middle)<br />

and a fragment <strong>of</strong> a trapezoidal-sectioned Green “type IV” adze from Zangana.<br />

404


Figure 110: Images <strong>of</strong> adzes illustrated on previous page.<br />

405


Figure 111: Canarium hammerstone from H5 ceramic findspot (top) (a similar artefact<br />

was found at Hoghoi); waisted sandstone slab from Zangana (middle); and chert flakes<br />

from Hoghoi (bottom).<br />

406


Table 43: Petrographic descriptions <strong>of</strong> lithic artefacts.<br />

Artefact ID Sit<br />

e<br />

B1.T50 10-<br />

15<br />

Unit Description<br />

8 13 In hand specimen a pale green-grey fine-grained adze fragment. In thin section a fine-grained sedimentary or volcanic<br />

rock. It is impossible to be more determinate without further analysis, as recrystallization <strong>of</strong> sedimentary clasts and<br />

primary volcanic crystallization could have the same petrographic result.<br />

HG.17.1 2 17 In hand specimen a glossy black stone, being a small fragment <strong>of</strong> an adze with polishing visible on two faces. In thin<br />

section this shows a fine-grained texture with sedimentary banding, with a possibility also <strong>of</strong> recrystallization.<br />

HG.99 2 99 A complete adze in a glossy black stone, similar in appearance to HG.17.1. This adze is reminiscent <strong>of</strong> Green and<br />

Davidson’s type V, occurring only at the early end <strong>of</strong> the Samoan sequence [Green, 1969 #375:25] In thin section,<br />

sedimentary or volcanic: comprising fine material in a very fine-grained matrix, with some evidence <strong>of</strong> feldspar<br />

recrystallization, with clay-rich silty minerals perhaps present as well.<br />

B3.T10 5-10 8 89 Mid-grey fine-grained appearance in hand specimen, weathered to a green-grey on the surface. In thin section coarser<br />

grained with micaceous clay precursors, dominated by plagioclase, with some Pyroxenes.<br />

A.12 8 48 In hand specimen a dark grey rock with a dull but generally unweatherd surface, being the sharpened end <strong>of</strong> a thin flat<br />

adze. In thin section a sedimentary rock, coarser grained. diagenesis indications include feldspar recrystallization and<br />

clay precursors. <strong>The</strong>re is no obvious calcite, and the matrix is dominated by plagioclase feldspars. Noted for<br />

recrystallized opaque pyrite.<br />

A.31 8 94 In hand specimen a complete but heavily resharpened plano-convex “horseho<strong>of</strong>”(Green and Davidson type 5) or<br />

gouge-edged adze in a glossy brown-grey stone (mid-grey when sawn). In thin section a coarser-grained sedimentary<br />

rock, similar to A.12., with feldspar dominant, diagenetic recrystallization <strong>of</strong> a variety <strong>of</strong> minerals, including opaque<br />

pyrite.<br />

A.6 8 23 In hand specimen the midsection <strong>of</strong> an adze, probably plano-convex, with surface weathered to a dark brown-grey,<br />

and mid-grey when sawn. In thin section the same as A.31 and A.12.<br />

B1.T50 0-<br />

5m<br />

8 12 A small grey rock fragment without obvious polished or ground surfaces. Mid-to-dark-grey when sawn. In thinsection<br />

it is unclear whether this is sedimentary or volcanic in origin, due to an advanced stage <strong>of</strong> recrystallization,<br />

making it difficult to see original textures. Dominated by plagioclase feldspar in various stages <strong>of</strong> recrystallization.<br />

Fine material also looks feldspathic, but includes a pyroxene/Ferromagnesian component. Pale-green chlorite is<br />

possibly present. Texture possibly indicates a plutonic origin, but this is tentative, and would require further analysis<br />

and thought to characterize this with more confidence.<br />

Miho 3 1 butt-end <strong>of</strong> a triangular-sectioned or fractured adze? Lightly ground or water rolled? Surfaces are dark-brown-grey,<br />

with sawn surfaces being mid grey to dark grey. This rock is a lot fresher than “B1.T50 0-5", with only slight<br />

recrystallization. Fine-grained plagioclase crystals are set in an even finer-grained matrix <strong>of</strong> recrystallized feldspathic<br />

character. This might be an immature, poorly-sorted sedimentary rock, or a crystal-rich igneous rock with a low<br />

carrying fluid content. XRF analysis would clarify whether chemistry indicates volcanic or sedimentary origin and<br />

diagenesis. X-ray diffraction would show up any quartz, clay minerals or calcite.<br />

P199 1 1 In hand specimen a thin small flake <strong>of</strong> fine-grained stone weathered to a golden-brown color, with light-grey sawn<br />

color. Possibly reworking debitage? In thin section this has a very fine-grained volcanic matrix carrying scattered<br />

plagioclase and pyroxene phenocrysts.<br />

HV.04.141 4 4 A flat sandstone slab 10x7x1cm, weathered to a golden-brown color on the surface and light grey when sawn. In thin<br />

section this is well supplied with pyroxenes, plagioclase and quartz, these being immature, angular, poorly sorted,<br />

with a silty clay cement. No Calcite. This is a packed stone with larger grains forming a self-supporting framework.<br />

Clinopyroxene/augite are present(a green pyroxene is probably augite). Plagioclase can show simple and multiple<br />

twinning as well as zoning. Amphibole and hornblende are present as well as a fine-grained foraminiferous silt clast.<br />

B(Zangana) 8 B In hand specimen this was a fin-shaped sherd <strong>of</strong> sandstone weathered to a golden brown, about 12cm maximum<br />

dimension, with a light grey color when sawn. In thin section a predominantly calcite cement with occasional<br />

unfragmented foraminiferal shell, suggesting calcite is a direct perecipitant, or a biogenetic calcite fine fraction, or<br />

recrystallization. <strong>The</strong> matrix is dominated by silt-sized glass fragments and some crystal products( feldspars including<br />

plagioclase, pyroxenes, and opaques). This was interpreted a s a cemented volcanic ash dominated by glassy<br />

fragments, with crystals co-genetic with the glass, which is bubble-fractured by expanding gasses. Microprobing the<br />

feldspars and glass or grinding/acid-washing/XRF would identify the type <strong>of</strong> volcanism involved.<br />

Z76 8 76 A coarse grey sandstone slab with pedestalled black mineral grains, 16x12x2cm and subrectangular in form, with two<br />

notches forming a waist, interpreted as a sandstone grindstone modified for secondary use as an anchor stone. Sawn<br />

colour matches the surface colour. In thin section fill <strong>of</strong> immature, angular, unsorted pyroxene, plagioclase and augite<br />

up to 1mm in length in a calcite cement. Fine-grained silty volcanogenic clasts are also present, which sometimes<br />

display fragments <strong>of</strong> coarser pyroxene and plagioclase. No confirmed quartz.<br />

It seems clear there are diverse resources being used over the period <strong>of</strong> deposition <strong>of</strong> these<br />

materials. Additional samples, more concerted analysis, and discovery and<br />

407


characterization <strong>of</strong> source quarries are needed to explicate patterns <strong>of</strong> lithic raw material<br />

transport. Adzes/fragments from Zangana <strong>of</strong> recrystallized fine-grained rock showed a<br />

variety <strong>of</strong> surface colours, including a pale greenish example. This suggests that raw<br />

material procurement may have been influenced by surface color, as there is some<br />

congruence here with the green adze rocks noted by Green for the Reef-Santa Cruz Lapita<br />

sites, the green plano-convex example from Samoa (Fijian origin?), the adze in green<br />

alteration dacitic welded tuff from Tonga, and green meta-sedimentary adzes from Naigani<br />

(Green 1979, 1994).<br />

A number <strong>of</strong> adzes and axes from Paniavile held in the private collection <strong>of</strong> Mr Sae<br />

Oka were photographed during the course <strong>of</strong> the study (Figure 112). Some small waisted<br />

axes and one unfinished adze preform were reportedly collected from Gharanga, by the late<br />

Mr Phillip Lanni (Figure 113, Figure 114). <strong>The</strong> preform is much longer than the complete<br />

adzes, while cross-section size is similar to the other adzes. If the preform is any indication,<br />

the adzes collected are less than half their original length, suggesting a lengthy use life<br />

involving a massive grinding resharpening labour or possibly more rapid<br />

reflaking/resharpening to repair edge damage.<br />

Sandstone artefacts/manuports similarly suggest a diversity <strong>of</strong> sources, as yet<br />

poorly represented in the collections. <strong>The</strong>re is clearly vast scope for artefact studies and<br />

lithic sourcing research for this type <strong>of</strong> intertidal ceramic/lithic scatter. HV.04.141 (Table<br />

43) is unlikely to be from the New Georgia group due to the abundance <strong>of</strong> quartz in the<br />

sand-sized component, as sediments rich in quartz should not occur given the geological<br />

characteristics <strong>of</strong> the area, although some forearc sediments have a plutonic component<br />

which might be relevant here (the nearest such source would be southern Rendova or<br />

Tetepare). <strong>The</strong> cemented volcanic ash (sample B-Zangana South, Table 43) might<br />

originate within the New Georgia group, as rocks originating as volcanic ashes are present<br />

on Ranongga, and possibly Southern Rendova or Tetepare.<br />

408


Figure 112: Artefacts photographed from Oka collection and reported to be from the<br />

Paniavile site: shell and stone adzes (top); stone adzes (middle) and waisted<br />

tools/weapons and a pineapple club (bottom).<br />

409


Figure 113: Waisted axes photographed from the Lanni collection, courtesy <strong>of</strong> the late<br />

Mr. Phillip Lanni, found in the vicinity <strong>of</strong> Gharanga Stream.<br />

410


Figure 114: Un-ground adze preform photographed from Lanni collection<br />

courtesy <strong>of</strong> the late Mr. Phillip Lanni, reportedly found at the Gharanga<br />

site.<br />

Sample Z76, from Zangana (the waisted sandstone slab in Figure 111), could conceivably<br />

be from Ranongga/Rendova/Tetepare, as the forearc region has had a complex history <strong>of</strong><br />

volcanism, uplift <strong>of</strong> the forearc region followed by subsidence, associated with erosional<br />

features, and subsequent rapid and accelerating uplift in response to subduction <strong>of</strong> the<br />

Coleman seamount (Mann et al. 1998). <strong>The</strong>se processes have created a variety <strong>of</strong><br />

siltstones, mudstones and sandstones <strong>of</strong> the Tetepare Formation.<br />

Chert Flakes/Fragments:<br />

Two chert flakes were found at Hoghoi (Figure 111). One (HG.20.22) was thin sectioned,<br />

411


and identified as microquartz 2 , but did not have any micr<strong>of</strong>oraminifera that might narrow<br />

the range <strong>of</strong> options for a source region. It is likely to be exotic to the New Georgia group<br />

as cherts are not expected there. <strong>The</strong> nearest known sources <strong>of</strong> chert are in Vella Lavella,<br />

where it occurs in minor amounts (Sheppard, pers. comm.), Santa Ysabel, Nggela or<br />

Malaita (Sheppard 1993:124).<br />

Analysis <strong>of</strong> Hoghoi Water-rounded and Fractured Volcanic Manuports:<br />

A transect-collected assemblage <strong>of</strong> 345 lithic manuports from the Hoghoi site was<br />

described in the field by matching to a type series <strong>of</strong> 23 stones or stone fragments. <strong>The</strong><br />

characteristics <strong>of</strong> degree <strong>of</strong> fragmentation (water rounded, fragmented with rounding, or<br />

fragment with no water-rounded cortex visible), type, and mass (to the nearest 50g) were<br />

recorded for each stone. <strong>The</strong>se data are given in table “Hoghoi_lithics.db” appended on<br />

CD. <strong>The</strong> type series was retained and petrographic thin-sections were prepared for each<br />

type. A combination <strong>of</strong> physical properties in hand-specimen and optical petrographic<br />

properties were used to reclassify the type series as shown in Table 44.<br />

Porphyritic olivine basalt is the most widespread rock type in the New Georgia<br />

group and occurs on all <strong>of</strong> the main islands except Simbo (Dunkley 1986:13). Classes 1,<br />

2, 10, 12, 16 and 18 fall into this category. <strong>The</strong> nearest source is Rendova, lying about<br />

12km distant across the Blanche channel, and this is a major source <strong>of</strong> ovenstones for the<br />

Roviana lagoon area in modern times. A slightly longer but more sheltered trip 15km east<br />

<strong>of</strong> Hoghoi can locate similar stones in mainland riverbeds draining into the Lagoon.<br />

Class 15 is a feldspar-phyric basalt or andesite, another important rock type that<br />

covers large areas <strong>of</strong> New Georgia, Vangunu, and southern Rendova. <strong>The</strong> nearest source<br />

<strong>of</strong> this rock type to the Hoghoi site is the Hura River (Dunkley 1986:22), about five km<br />

distant across the Lagoon. Vesicular and scoracious lava flows are present in the Picritic<br />

lavas eroding into the Viru Harbour area and Marovo Lagoon (Dunkley 1986:16), and<br />

2 Cherts were examined petrographically by Dr Peter Sheppard <strong>of</strong> the<br />

Anthropology Department, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

412


Classes 4 and 5 might be from such a source, which would involve transport from Viru<br />

Harbour at a minimum, a distance <strong>of</strong> about 50km. Kolombangara, at a similar distance<br />

from Hoghoi, is also noted by Dunkerly as having scoria cones. Closer sources are not<br />

precluded on current evidence. Class 6 has similar mineral composition to Class 5, and may<br />

be from a similar source.<br />

Class 7 is an olivine basalt with olivines serpentinised rather than altered to<br />

iddingsite or magnetite, suggesting a magnesium-rich magma. Sepentinised olivines are<br />

noted in the gabbros <strong>of</strong> the Choe intrusive complex (Dunkley 1986:31), but whether this<br />

means volcanic rocks in that area are more likely to show serpentine rather than iddingsite<br />

as the alteration product <strong>of</strong> olivine is unknown, but this suggests a specific question for<br />

source sampling in the future. If it does turn out that serpentine is characteristic <strong>of</strong> Viru<br />

harbour rocks, then the Class 4 and 5 scoriacious manuports, Class 6, and Class 7 could<br />

all be a pointer to procurement <strong>of</strong> a minority <strong>of</strong> manuports at a greater distance than is the<br />

norm today. Scoracious rocks in particular might have a non-cooking function, as their<br />

heat capacity is low. That function could relate to their relatively low specific gravity, and<br />

surface roughness which might confer desirable properties as net weights or possibly<br />

abrasives.<br />

Looking only at those manuports that were water-rounded and unfractured in form,<br />

a comparison <strong>of</strong> size (average weight) by class is given in Table 45. Some classes appear<br />

mainly as small cobbles or pebbles <strong>of</strong> less than 120g average weight (for example the<br />

large Class 18, and also classes 3, 5, 7, 10 and 16), while other classes average more than<br />

240g (Classes 1 and 2). This is most likely a slight difference in the survival rate <strong>of</strong> larger<br />

stones if used for cooking (i.e. Class 18 is more likely to shatter than Class 1) but the<br />

larger number <strong>of</strong> complete water rounded stones in Class 18 is a bit <strong>of</strong> a mystery, unless<br />

these were preferred for some other function, such as fishing or net weights. Perhaps this<br />

is just sampling error, as the number <strong>of</strong> occurrences <strong>of</strong> complete water-rounded stones is<br />

low for each class.<br />

413


Table 44: Petrographic classification <strong>of</strong> manuports collected at Hoghoi.<br />

Sampl<br />

e #<br />

count class Description<br />

HG1 67 1 Medium grey color, phenocrysts visible to the naked eye. In thin section a groundmass <strong>of</strong> Plagioclase laths and fine opaques<br />

with Pyroxene phenocrysts and a brown-rimmed alteration mineral (probably Iddingsite).<br />

HG2 36 2 In hand specimen a pinkish rock with pale green phenocrysts. In thin section the groundmass was predominantly Plagioclase<br />

laths with subordinate Olivines altered to Iddingsite. Phenocrysts were large Pyroxenes. No Feldspar phenocrysts.<br />

HG3 21 3 Fine grained spherulitic texture, a brown glassy matrix carrying scattered large Feldspar phenocrysts as well as Feldspar<br />

elongate quench crystals, giving stong indications <strong>of</strong> rapid cooling forming glass followed by recrystallization from a series<br />

<strong>of</strong> points. A large crystal <strong>of</strong> Calcite is present, pseudomorphing pre existing Olivine or Pyroxene.<br />

HG4 21 4 Pinkish in hand specimen with small vesicles. In thin section a groundmass <strong>of</strong> medium-textured Plagioclase and Pyroxene<br />

crystals.<br />

HG5 26 5 In hand specimen pinkish, fine-grained, s<strong>of</strong>t to saw. A dense, fine-grained texture with some vesicularity. Reddish<br />

groundmass, with mainly Plagioclase phenocrysts and some Pyroxene.<br />

HG6 1 6 In hand specimen pale grey to pale green and fine-grained in hand specimen. In thin section this had a very fine-grained<br />

Plagioclase groundmass with Plagioclase phenocrysts, similar to HG5 but without the vesicular texture.<br />

HG7 10 7 In hand specimen grey, medium grained, even textured. In thin section a plagioclase groundmass and abundant pyroxene<br />

phenocrysts were noted, and also an alteration product <strong>of</strong> Olivine, Serpentine, indicating a relatively magnesium-rich<br />

magma.<br />

HG8 2 1 A coarse-grained grey rock in hand specimen, and in thin section a coarse volcanic Plagioclase groundmass with abundant<br />

Pyroxene phenocrysts and smaller brown alteration mineral, possibly Iddingsite.<br />

HG9 13 9 Grey and medium-textured in hand specimen. In thin section a very fine groundmass <strong>of</strong> unidentified mineral grains with<br />

large phenocrysts <strong>of</strong> Plagioclase and smaller Pyroxene phenocrysts.<br />

HG10 18 10 In hand specimen a dark-grey rock with greenish phenocrysts. In thin section a fine grained volcanic matrix, some small<br />

crystals <strong>of</strong> an alteration mineral, possibly Iddingsite, with phenocrysts <strong>of</strong> Pyroxene and Plagioclase.<br />

HG11 15 5 In hand specimen a s<strong>of</strong>t pinkish fine-grained rock, and in thin section a dense, fine-grained texture carrying large phenocrysts<br />

<strong>of</strong> Plagioclase and Pyroxenes, with a possible Olivine alteration product in the groundmass.<br />

HG12 12 12 In thin section a coarse-grained phenocrystic volcanic carrying Clinopyroxene, Augite, occasional Orthopyroxene<br />

(Hypersthene), red Iddingsite. (This alteration product <strong>of</strong> Olivines occurs in Iron-rich magmas and could result from use <strong>of</strong><br />

heat as oven-stones).<br />

HG13 6 5 In hand specimen s<strong>of</strong>t, pinkish and fine grained. In thin section a very fine-grained volcanic groundmass with mostly<br />

Plagioclase phenocrysts.<br />

HG14 4 14 A fine-grained light-grey water-rounded pebble. In thin section there is dense accumulation <strong>of</strong> Plagioclase and Pyroxenes in<br />

layers, could be plutonic. No Olivine. Fine-grained matrix suggests volcanic.<br />

HG15 10 15 Huge Feldspar/Pyroxene phenocrysts (some 1cm) in a matrix <strong>of</strong> fine Plagioclase. Some brown-rimmed smaller phenocrysts<br />

are possibly an altered pre-existing Olivine.<br />

HG16 14 16 In hand specimen a dark grey medium-grained rock similar to HG7. In thin section a matrix <strong>of</strong> very fine Plagioclase contains<br />

subordinate medium-sized Plagioclase phenocrysts and dominant large Pyroxene phenocrysts.<br />

HG17 7 17 In hand specimen a very-fine-grained dark grey angular fragment, in thin section a very fine grained volcanic matrix carrying<br />

scattered phenocrysts <strong>of</strong> Plagioclase and a dark brown/red alteration mineral, with occasional Pyroxene phenocrysts (see also<br />

GW8 in ?), interpreted as a fragment <strong>of</strong> adze rock.<br />

HG18 57 18 Dark grey in hand specimen with greenish phenocrysts, similar to HG 7 and HG 16. In thin section a medium-textured<br />

Plagioclase groundmass with Pyroxene phenocrysts and a lot <strong>of</strong> medium-sized opaques. Some brownish minerals that are<br />

possibly an alteration mineral <strong>of</strong> Olivine or Calcite.<br />

HG19 6 19 Vesicular in hand specimen with pink-to grey color zonation (heat alteration rom use?). In thin section a very fine<br />

groundmass (mineral not identified) with rounded red-to-brown phenocrysts filling some vesicles.<br />

HG20 4 2 In hand specimen dark grey and coarse-grained, and in thin section a coarse-grained texture, predominantly plagioclase<br />

groundmass with large Pyroxene phenocrysts. Large plagioclase is absent. Iddingsite phenocrysts present<br />

HG<br />

“incise<br />

d”<br />

pebble<br />

HG<br />

40-45<br />

pound<br />

er<br />

1 3 As below, except texture is slightly coarser and includes pyroxene phenocrysts.<br />

1 3 In thin section a fine grained sperulitic texture, a glassy matrix carrying scattered large Feldspar phenocrysts as well as<br />

Feldspar elongate quench crystals, giving strong indications <strong>of</strong> rapid cooling forming glass followed by recrystallization from<br />

a series <strong>of</strong> points.<br />

414


Table 45: Size comparisons between petrographic classes <strong>of</strong> lithic manuports.<br />

Class Count <strong>of</strong> class Sum <strong>of</strong> mass (g) Mean(St.D.) mass (g)<br />

1 15. 3600. 240(252)<br />

2 10. 2600. 260(353)<br />

3 17 1300 76(26)<br />

4 2. 300. 150(71)<br />

5 8. 825. 103(123)<br />

7 4. 400. 100(71)<br />

10 7. 550. 79(27)<br />

14 3. 575. 192(267)<br />

16 9. 1075. 119(146)<br />

18 31. 3650. 118(188)<br />

Taking water-rounded (complete) cobbles <strong>of</strong> all classes together (Figure 115), there is a<br />

suggestion in a histogram <strong>of</strong> sizes that stone collecting strategies operated between the<br />

limits <strong>of</strong> roughly 50g ( the scale used was not very accurate) and 1150g, although the<br />

lower limit may reflect an archaeological collection threshold, and the upper limit is<br />

imprecise due to small sample size <strong>of</strong> larger complete stones (heavier stones than this<br />

were present among the fractured cobbles, with the largest stone being 1900g, with mean<br />

mass for all stones collected being 200g, and Standard deviation <strong>of</strong> 212g). Most <strong>of</strong> the<br />

larger stones within these limits would have been fractured by heat, but the smaller stones,<br />

from 50g to 150g, seem to have been either not subjected to heating or have survived<br />

better. <strong>The</strong>se smaller stones are an inconvenient size for domestic cooking from the<br />

perspective <strong>of</strong> the modern local practice <strong>of</strong> handling stones individually with bamboo<br />

tongs, and may have had another function, such as net weights or fishing weights, for large<br />

nets such as turtle or dugong nets, which required heavier weights due to their added<br />

buoyancy.<br />

415


Figure 115: Water-rounded lithic manuports, cortex<br />

complete.<br />

Figure 116: Size distribution <strong>of</strong> fractured lithic manuports,<br />

either with some cortex or without.<br />

416


Ethnographic cooking practice need not reflect those <strong>of</strong> 3000 years ago though.<br />

If the weight frequency distribution <strong>of</strong> complete stones (Figure 115) is contrasted<br />

with the weights <strong>of</strong> fractured stones (Figure 116) it seems that the complete recovered<br />

examples underestimate the upper weight limit <strong>of</strong> procurement, as some fractured stones<br />

approach 2kg in weight.<br />

<strong>The</strong> spatial distribution at Hoghoi <strong>of</strong> the petrographic classes defined in this section<br />

is discussed in Chapter 11.<br />

Chapter Summary:<br />

Lithic manuports systematically surface collected at Hoghoi included a diversity <strong>of</strong> volcanic<br />

rocks, with coarse porphyritic textures and cortex suggestive <strong>of</strong> origins as transported<br />

water-rounded cobbles. Sizes in excess <strong>of</strong> 2kg were targeted in some cases, but there was<br />

no evidence for any rocks larger than this. <strong>The</strong>se coarse textures, and the fractured forms<br />

<strong>of</strong> most rocks suggests a cooking function. Some fine-grained sperulitic-textured<br />

fragments with quench-textures suggest rocks poorly suited to oven-stone use (liable to<br />

explode?), and some <strong>of</strong> these may be fragments <strong>of</strong> flaked artefacts. At least one fine-<br />

grained recrystallysed sedimentary or volcanic fragment in the manuport collection is likely<br />

to be a fragment <strong>of</strong> an adze, as petrographic characteristics matched those <strong>of</strong> some <strong>of</strong> the<br />

adzes collected.<br />

A number <strong>of</strong> smaller porphyritic water-rounded pebbles were collected, which do<br />

not fit with local ethnographically observed oven-stone size selection. <strong>The</strong>se may be<br />

evidence <strong>of</strong> other non-cooking functions, such as net weights or fishing sinkers, or possibly<br />

slingstones even, or may indicate a shift in cooking practices over time. <strong>The</strong> presence <strong>of</strong><br />

scoracious basalts is inconsistent with a cooking function also, and these may originate<br />

from Viru or Kolombangara, although more local sources may exist, as yet<br />

417


unreported. While most olivine basalts are most likely from Rendova, serpentinised olivines<br />

may indicate different sources, but this has yet to be tested.<br />

A diversity <strong>of</strong> adze materials indicated the potential for sourcing studies to<br />

illuminate raw material procurement patterns for this site type. At least one plano-convex<br />

horseho<strong>of</strong>-edged adze form was recovered from Zangana, regarded by Green as indicating<br />

a homologous link with Lapita at this site, as these forms are characteristic <strong>of</strong> the Lapita<br />

horizon and derivatives (Green, pers. comm.). Comparison <strong>of</strong> the complete adzes with an<br />

unused preform suggests the complete used examples have been heavily resharpened to<br />

less than half their original length prior to discard, suggesting little remaining use life.<br />

Although the sample <strong>of</strong> such items is small, it does suggest accumulation from gradual<br />

refuse discard rather than discard as a result <strong>of</strong> sudden or catastrophic abandonment <strong>of</strong><br />

sites or extinction <strong>of</strong> inhabitants. <strong>The</strong> presence <strong>of</strong> a number <strong>of</strong> small fragments <strong>of</strong> adzes,<br />

identifiable only through raw material characteristics, suggests either a taphonomic regime<br />

capable <strong>of</strong> smashing these rugged artefacts, or reworking <strong>of</strong> adzes into smaller artifacts,<br />

damning any notion that the observed quantities <strong>of</strong> even rugged lithic artefacts can be read<br />

as directly measuring the intensity/duration <strong>of</strong> discard.<br />

Other lithic items such as sandstone fragments and artefacts are not always easy to<br />

find among coral gravels, but the diversity <strong>of</strong> sources indicated by petrographic analysis<br />

suggests these can be a major component <strong>of</strong> sourcing studies, and are worth putting more<br />

effort into finding and analysing.<br />

This <strong>chapter</strong> demonstrates that archaeological values <strong>of</strong> this type <strong>of</strong> site lie as<br />

much in the tiny nondescript lithic fragments as in the complete artefacts. One is <strong>of</strong> little<br />

value without the other. <strong>The</strong> complete artefacts alone would mislead if read as the<br />

quantities and diversity <strong>of</strong> adzes or whatever discarded in the past. Similarly, the<br />

fragments alone would be difficult to recognize as artefact fragments without petrographic<br />

analysis <strong>of</strong> both these and the complete or identifiable adzes. Intensive investigation <strong>of</strong><br />

lithic variability can unlock the cultural and postdepositional formation processes <strong>of</strong> these<br />

418


sites. It is clear from this lithic analysis, as it was from the various ceramic <strong>chapter</strong>s, that<br />

the information content <strong>of</strong> these sites is fragile as a result <strong>of</strong> this interdependence: take the<br />

fancy bits away and it gets harder, if not impossible to understand the fragments, and the<br />

fancy pieces alone do not give an accurate picture <strong>of</strong> site formation.<br />

<strong>The</strong> raw materials and forms <strong>of</strong> the Roviana flaked and polished stone adzes<br />

recovered from the intertidal sites are very similar to the materials recovered from the<br />

Buka Lapita sites on the reef flat. This similarity is so marked as to suggest that detailed<br />

petrographic and morphological comparisons would be justified, although it was not<br />

possible to do this within the time-constraints <strong>of</strong> the present study.<br />

Comparisons between the Buka sample and the Roviana sample <strong>of</strong> adzes suggest<br />

these are expressions <strong>of</strong> a very similar range <strong>of</strong> adze forms and selection <strong>of</strong> raw materials,<br />

most economically explained as homologous similarity. This supports the notion <strong>of</strong> a<br />

Lapita/post-Lapita stone adze repertoire, if not an adze kit, which is not to say that there<br />

was no change over time or space, but which acknowledges that given the limited current<br />

sample, particularly for near-Oceania, a coarse stone adze horizon can be seen, which<br />

probably occupies a broader temporal span than elaborate dentate-stamped pottery, and<br />

serves to link the “post-Lapita” Roviana intertidal sites to the Lapita ceramic series by<br />

homologous similarity <strong>of</strong> the lithic adze component <strong>of</strong> sites.<br />

<strong>The</strong> near-Oceania Lapita-horizon stone adze sample is now dominated by the items<br />

recovered from the Buka/Nissan reef sites and the Roviana intertidal sites. This suggests<br />

that further sampling <strong>of</strong> this settlement type holds the prospect <strong>of</strong> major dividends for this<br />

rare class <strong>of</strong> artefact, and that a large sample attained by this means can be expected to<br />

enable detailed sourcing and variability studies to proceed in the future, to an extent never<br />

before possible in Near-Oceania due to sample size constraints.<br />

419


420


Introduction:<br />

CHAPTER 11:<br />

INTRASITE SPATIAL STRUCTURE OF<br />

CERAMIC AND LITHIC VARIABILITY:<br />

Intrasite spatial structure is here defined as the patterned distribution <strong>of</strong> potsherds,<br />

artefacts, and/or manuports across sites (Wandsnider 1996). Intrasite spatial structure in<br />

the Roviana intertidal scatters is assumed to result from processes such as patterned<br />

discard (which may or may not have included contemporaneous differentiated activity<br />

areas), wave sorting <strong>of</strong> archaeological materials, selective trapping and preservation <strong>of</strong><br />

archaeological materials by site micro-topography and/or postdepositional pattern-<br />

alteration by humans.<br />

Site occupation span can be expected to have varied from site to site, and sites in<br />

favourable locations may have been occupied more than once. Even within a continuous<br />

occupation, it seems likely that the extent and locus <strong>of</strong> settlement may have changed over<br />

time, the latter either randomly, or possibly through unknown socially-determined<br />

mechanisms such as the avoidance <strong>of</strong> places formerly used by the deceased. Such processes<br />

should be expected to create, over the generations, chrono/stylistic spatial structure within<br />

sites. Ultimately the process <strong>of</strong> distinguishing whether stylistic spatial structure represents<br />

chronological change in style or contemporaneous diversity is one that occurs in dialogue<br />

with one’s ideas on ceramic classification and the nature <strong>of</strong> the stylistic series (Baxter<br />

1994:23). For a poorly-known sequence like the Roviana one, with no well-established<br />

fine-grained stylistic chronology, any conclusions must be regarded as preliminary working<br />

hypotheses.<br />

421


Review <strong>of</strong> Methods <strong>of</strong> Spatial Analysis:<br />

While intersite distributional mapping <strong>of</strong> ceramic styles or types within a geographic<br />

approach was common in European archaeology (e.g. Hodder & Orton 1976)and has been<br />

applied in the Pacific (Irwin 1985), ceramic intrasite spatial studies have been relatively<br />

rare in archaeology (Fontana 1998) compared to intersite or distributional studies, many<br />

examples being student dissertations ( e.g. Henderson 1998, Henderson 1992, Robertson<br />

2001). <strong>The</strong>re are notable exceptions to this tendency, some local (Best 1984:594-617,<br />

Cowgill et al. 1984, Fontana 1998, Sheppard & Green 1991).<br />

During the 1980s, debate in intrasite spatial analysis centred around spatial<br />

clustering and the methods by which archaeologists could identify spatially clustered<br />

patterns (Baxter 1994:23-24). LuAnn Wandsnider noted that emphasis has shifted in recent<br />

studies from a functional view <strong>of</strong> site structure in which activity patterning ought to be<br />

easily read and toolkits easily identified spatially (e.g Blankholm 1991), to a broader and<br />

more complex formational view <strong>of</strong> spatial structure, where the interaction <strong>of</strong> both cultural<br />

and natural formation processes are acknowledged as responsible for the structure <strong>of</strong><br />

archaeological deposits, and the identification <strong>of</strong> specific formational processes has become<br />

the goal (Wandsnider 1996). Congruent with this shift has been a shift in method from<br />

partitive typological approaches seeking to characterized artifact clusters within sites as<br />

functionally discrete to a more inclusive set <strong>of</strong> methods that examine distributions <strong>of</strong> a suite<br />

<strong>of</strong> formationally sensitive attributes <strong>of</strong> artefacts across sites. Although Wandsnider does<br />

not specifically make this point, it appears from the historical data presented that a shift in<br />

technique <strong>of</strong> analysis accompanied this trend, from computer identification <strong>of</strong> spatial<br />

clusters, to a recent state <strong>of</strong> practice in which it is more common to rely on visual<br />

inspection <strong>of</strong> density maps <strong>of</strong> formationally sensitive indicators (Wandsnider 1996:Table<br />

1).<br />

An analysis <strong>of</strong> intrasite stylistic spatial structure can be used to make inferences<br />

about changes in the extent/location <strong>of</strong> settlements over time. Most open sites are<br />

palimpsests <strong>of</strong> formation events and processes: in relation to intrasite patterning generally,<br />

422


Kintigh describes the problem <strong>of</strong> overlap or mixing <strong>of</strong> deposits as “perhaps the outstanding<br />

issue before (spatial archaeology)” (Kintigh 1990), and I suggest that the prospects for<br />

advance in this respect are best for the special case <strong>of</strong> ceramic archaeology, where<br />

relatively rich stylistic information can be expected to be a sensitive chronological<br />

indicator.<br />

Selection <strong>of</strong> Sites for Intrasite Spatial Analysis:<br />

Entire-sites were collected as single units during initial intertidal surface collections in early<br />

1996. By 1997, when intertidal survey was concentrated in the Kaliquongu region <strong>of</strong><br />

Roviana Lagoon, the aim was to try to document intrasite spatial structure for sites. Hence,<br />

the sites Zangana, Gharanga (2 nd collection), Honiavasa and Hoghoi (2 nd collection) were<br />

surface-collected in multiple transects. <strong>The</strong> exception was the Miho site, which was <strong>of</strong><br />

more limited extent, where significant along-shore spatial structure seemed unlikely.<br />

Intertidal ceramic/lithic scatters at H5, H6 and H7 were not surface collected due to time<br />

constraints and the small samples <strong>of</strong> sherds present. Largely aceramic intertidal lithic sites<br />

at Mbolave Island, H4, H7 and at Elelo/Mbanga were left undisturbed for the present.<br />

Objectives:<br />

Analysis <strong>of</strong> vessel completeness (Chapter 5) showed that only a small percentage <strong>of</strong> the<br />

deposited pottery was recovered, suggesting a harsh taphonomic regime for most sherds,<br />

resulting in the complete destruction <strong>of</strong> most vessels, and lucky survival in good condition<br />

for a minority <strong>of</strong> sherds from a minority <strong>of</strong> vessels. It seemed unlikely therefore that<br />

spatial distribution <strong>of</strong> sherds corresponding directly with activity areas (Schiffer 1995b)<br />

would have been preserved, although the possibility was not ruled out. Of more direct<br />

interest though, is the question <strong>of</strong> whether there were any broad-scale differences within<br />

423


sites in the distribution <strong>of</strong> fabric or style. Such differences might be expected to survive<br />

the taphonomic regime, and would be significant to chronological analysis <strong>of</strong> style, as sites<br />

could potentially be split into smaller seriation units on the basis <strong>of</strong> such distributional<br />

patterning, potentially lessening the problem <strong>of</strong> chronological mixing <strong>of</strong> styles within site<br />

assemblages.<br />

If wave-sorting <strong>of</strong> sherds has been a primary factor in creating spatial structure,<br />

small sherds should be found along the strandline, the upper intertidal, the swash zone,<br />

should be relatively free <strong>of</strong> sherds, and larger sherds should be found at the deeper margins<br />

<strong>of</strong> sites, in shallow water at low tide. <strong>The</strong> Zangana sherd size data was analysed in these<br />

terms. Furthermore, if wave-sorting <strong>of</strong> sites has been a primary factor in creating spatial<br />

structure, along-shore movement <strong>of</strong> smaller sherds would create a spread distribution <strong>of</strong><br />

small sherds in relation to large sherds, which latter should more nearly reflect the location<br />

and extent <strong>of</strong> primary discard. <strong>The</strong> along-shore linear nature <strong>of</strong> the Hoghoi ceramic scatter<br />

was considered an ideal vehicle for an analysis <strong>of</strong> this sort, where the narrow across-shore<br />

extent <strong>of</strong> the site precluded much in the way <strong>of</strong> analysis <strong>of</strong> across-shore sorting. For the<br />

Hoghoi site, lithic manuports were used in addition to ceramics to address these questions.<br />

<strong>The</strong> Honiavasa and Gharanga sites were collected in unequal-area transects. In the<br />

case <strong>of</strong> the Honiavasa site this was due to collection during a rising tide in deeper water<br />

and windy weather, where collection proceeded along five pace-measured transects<br />

between an anchored canoe and the shore, arranged radially in a fan formation around<br />

Honiavasa Point. <strong>The</strong> accuracy <strong>of</strong> measurement <strong>of</strong> area <strong>of</strong> the resulting transects is<br />

unknown, so relative measures <strong>of</strong> assemblage composition by transect are used wherever<br />

possible.<br />

One aim was to seek, through spatial structure, to untangle the sources <strong>of</strong><br />

decorative variability, and another was to take a distributional approach to sherd size,<br />

treating artefact fragments as sedimentary particles. While there may be some lag deposits<br />

<strong>of</strong> larger sherds that preserve behavioural or activity patterning, these also may be highly<br />

424


correlated with site microtopography and minimally correlated with specific activity areas.<br />

Method:<br />

A statistical approach was eschewed in favour <strong>of</strong> visual inspection <strong>of</strong> graduated symbol<br />

maps representing sherd counts or other quantification <strong>of</strong> cell compositional attributes,<br />

similar to the display used for ground visibility data in the Hvar Archaeological Survey in<br />

Yugoslavia (Gaffney et al. 1991). A difference from the Hvar displays that should be noted<br />

is that in the present study the dot diameter is related to the square root <strong>of</strong> the raw transect<br />

data, whereas in the Hvar survey a limited number <strong>of</strong> dot sizes was used to display data<br />

that had been converted to ranges or categories.<br />

An advantage <strong>of</strong> this graduated dot approach over isopleth displays is that the<br />

displayed data (symbol sizes) are derived fairly directly from the raw data (sherd counts<br />

or sums <strong>of</strong> sherd areas by cell), making a visual assessment <strong>of</strong> sample size effects by<br />

transect possible. An advantage <strong>of</strong> this type <strong>of</strong> display over choropleth displays was that<br />

the elongated transect shapes <strong>of</strong> the field collection strategy would have been distracting<br />

to the eye had these displays been used, whereas size-graduated circular symbols created<br />

a directionally-unbiased visual effect. Were frequency data or ratio data displayed instead,<br />

significant sample size information would be less accessible.<br />

Sherds were coded by numbered collection unit for intra-site spatial analysis in<br />

table “Flat.db” as follows:<br />

425


Paniavile (site 1): 1=cave assemblage collected prior to ‘89?<br />

2=Jan 1996 colln.<br />

Gharanga: (site 5) 1=Jan 96<br />

3=August 96 colln.<br />

2=G west <strong>of</strong> stream 98<br />

3=G East <strong>of</strong> stream 98<br />

Hoghoi (Site 2) 1-21=5m wide transects<br />

99=initial surface collection by Kenneth Roga August 97<br />

130=hg13t(subsurface test)<br />

150=hg15a(additional)<br />

151=hg15s(subsurface test)<br />

152=hg15t (another test)<br />

Honiavasa (Site 4) 1-5=radial transects 1-5<br />

Kopo (Site 7) 1=single collection unit<br />

Miho (Site 3) 1=single collection unit<br />

Nusa Roviana (Site 6) 1=single collection unit<br />

Zangana (Site 8) 1-105= transect units as per Figure 117(collected August97)<br />

110=initial surface collection (June97) <strong>of</strong> Zangana<br />

111=last collection (August 1997) <strong>of</strong> Zangana<br />

Digital site maps were drawn for Sites 2, 4 and 8 (Hoghoi, Honiavasa and Zangana) from<br />

field notes and tape/compass maps using AUTOCAD drafting s<strong>of</strong>tware. <strong>The</strong>se maps were<br />

linked to ceramic thematic data tables using ArcInfo and MapInfo programs. MapInfo was<br />

used to create site graduated dot thematic maps.<br />

426


Figure 117: Intertidal collection units at Zangana: numbers are values in “unit” column<br />

in table “Flat.db” appended on CD. Units without numbers are those which yielded no<br />

sherds.<br />

427


Zangana:<br />

Natural Formation Processes at Zangana:<br />

Across shore size sorting. <strong>The</strong> question: can I demonstrate that appreciable numbers <strong>of</strong><br />

sherds have moved, and if some sherds may not have moved much, can this be shown?<br />

It would be unsatisfactory to analyse the Zangana materials purely in term <strong>of</strong> distance from<br />

shore in relation to sherd size, for two reasons: firstly, the shoreline has undergone<br />

substantial modification during WWII as the site <strong>of</strong> a major American landing on New<br />

Georgia, and more recently during refurbishment <strong>of</strong> the road and wharf circa 1988;<br />

secondly, the topography <strong>of</strong> the intertidal zone varies significantly between the North and<br />

South <strong>of</strong> the site.<br />

Zangana Sherd Count<br />

All temper classes<br />

80<br />

40<br />

8<br />

New Georgia Mainland Roviana Lagoon<br />

428<br />

35m<br />

�<br />

Figure 118: Spatial distribution <strong>of</strong> the sherd sample at<br />

Zangana.


<strong>The</strong> dense cluster <strong>of</strong> ceramic sherds at the north end <strong>of</strong> the site (Figure 118) occur<br />

where the intertidal is characterized by relative microtopographic roughness, and overall<br />

narrowness <strong>of</strong> the intertidal and shallow subtidal (around 15m from the high water mark<br />

to the lower limit <strong>of</strong> the intertidal). This microtopographic roughness is thought to have<br />

acted as a sherd trap and shield, and supplies the primary explanation for the preservation<br />

<strong>of</strong> sherds in this area. By contrast, the intertidal zone south <strong>of</strong> the wharf is generally about<br />

30m wide during the low tides, with sherds tending to be found in the recesses <strong>of</strong> a more<br />

gentle microterrain.<br />

North <strong>of</strong> the Zangana wharf there are two clusters <strong>of</strong> transects with reasonable<br />

sample sizes, interspersed by an area <strong>of</strong> s<strong>of</strong>t sediment most likely deposited as a result <strong>of</strong><br />

the sheltered conditions from the SE tradewind and southerly sea breezes created by the<br />

earthen wharf. South <strong>of</strong> the wharf there is a band <strong>of</strong> transects along the shoreline which<br />

have marginal samples <strong>of</strong> around eight sherds each, and a zone or band towards the lower<br />

limit <strong>of</strong> the intertidal zone which includes some slightly larger samples, comprising three<br />

groups <strong>of</strong> transects.<br />

Zangana Average Sherd Area<br />

(cm2)<br />

110<br />

55<br />

11<br />

New Georgia Mainland Roviana Lagoon<br />

429<br />

35m<br />

�<br />

Figure 119: Across-shore size sorting at Zangana.


<strong>The</strong> larger sherds are along the seaward margins <strong>of</strong> the intertidal zone in most cases<br />

(Figure 119), suggesting selective removal <strong>of</strong> smaller sherds and probably also <strong>of</strong> sherds<br />

not fortuitously sheltered/trapped by microtopography. This pattern suggests the seaward<br />

margin <strong>of</strong> Zangana South is a lag deposit, while the landward margin is a secondary<br />

accumulation <strong>of</strong> wave-transported sherds. <strong>The</strong> large sherd counts from transects at the<br />

north end <strong>of</strong> the site comprise mainly small sherds, with a couple <strong>of</strong> larger averages<br />

occurring only in transects with small samples. <strong>The</strong>re is a suggestion though <strong>of</strong> occasional<br />

larger sherds around the seaward margins <strong>of</strong> this area also.<br />

Sherds immediately to the north <strong>of</strong> the wharf are all small, and it is here that WWII<br />

and subsequent effects are probably greatest, with trampling and recent sherd breakage<br />

likely to be extreme due to modern foot traffic; also, accumulation <strong>of</strong> s<strong>of</strong>t sediments<br />

following rebuilding <strong>of</strong> the wharf in the 1990s has reduced archaeological visibility here.<br />

<strong>The</strong> more extensive tidal flat to the south <strong>of</strong> the wharf displays a more coherent<br />

pattern <strong>of</strong> size sorting across the shoreline, with all <strong>of</strong> the strandline sherd samples having<br />

average sherd sizes <strong>of</strong> less than 8cm 2 . All this supports the intuitive impression formed<br />

during field collection that the strandline comprised almost exclusively small, water<br />

rounded sherds, while the seaward limits <strong>of</strong> the intertidal zone yielded larger sherds with<br />

little evidence <strong>of</strong> rolling damage, and only minor sand abrasion in many cases, marine<br />

growth probably being a factor in prevention <strong>of</strong> the latter.<br />

A further question: was there a smaller locus or multiple loci <strong>of</strong> settlement in the<br />

past that has been spread into the present 3000m 2 distribution by wave-transport <strong>of</strong><br />

sherds? It seems clear from the above that wave processes have created a pattern <strong>of</strong> small<br />

sherds along the strandline which are relatively mobile and larger sherds further seaward<br />

that are less so. Is there a spread <strong>of</strong> small sherds and a lag deposit <strong>of</strong> larger sherds? Or<br />

perhaps multiple spaced lag deposits <strong>of</strong> larger sherds interspersed with small sherds<br />

representing spread from multiple occupation loci? This latter possibility seems borne out<br />

to some extent by the distribution in Figure 119, particularly to the south <strong>of</strong> the wharf. A<br />

band <strong>of</strong> larger sherds is distributed sub-parallel to the shoreline, possibly following a pre-<br />

bulldozer shoreline around what would have been the less prominent prehistoric Zangana<br />

point. Whether clusters in this distribution are a random sample size effect or a<br />

taphonomic effect relating to microtopographic variation are questions for future research<br />

430


on these sites. A third explanation has not been ruled out, that these clusters <strong>of</strong> larger<br />

sherds forming a line along the shore are a settlement pattern, a line <strong>of</strong> discard zones<br />

surrounding house locations.<br />

If sherd size and quantity are combined by summing sherd areas by transect, giving<br />

weight to those transects having the greatest total surface area <strong>of</strong> pottery, a picture that<br />

could be interpreted in this way emerges (Figure 120). South <strong>of</strong> the wharf, there are a line<br />

<strong>of</strong> four holes in the distribution <strong>of</strong> potsherds which might result from a line <strong>of</strong> structures<br />

surrounded by throw zones for broken pottery. A test <strong>of</strong> this idea would be to examine<br />

whether these gaps could be random, and whether they correlate with microtopographic<br />

variation, but these tests will not be attempted here.<br />

Zangana Sum <strong>of</strong> Sherd Areas<br />

(cm2)<br />

620<br />

310<br />

62<br />

New Georgia Mainland Roviana Lagoon<br />

431<br />

35m<br />

�<br />

Figure 120: Possible linear settlement patterning in the<br />

distribution <strong>of</strong> large sherds at Zangana south.


If this distribution does represent a settlement pattern, this would imply either long-<br />

duration stability in settlement layout, or short duration <strong>of</strong> occupation, as, were one <strong>of</strong><br />

these conditions not the case, such patterning would be obscured by overprinting.<br />

Spatial Structure in Ceramic Style at Zangana:<br />

<strong>The</strong> above discussion established that post-depositional taphonomic effects have probably<br />

obscured most small-scale settlement patterning, with the possible exception <strong>of</strong> the<br />

Figure 121: Initial point-provenanced collection <strong>of</strong><br />

decorated sherds at Zangana (subsequent collection<br />

transects shown for spatial reference).<br />

southern-seaward part <strong>of</strong> the site. Initial surface collection <strong>of</strong> a small number <strong>of</strong> sherds at<br />

Zangana targeted only decorated sherds. Point locations <strong>of</strong> sherds were recorded.(Figure<br />

121)<br />

432


<strong>The</strong>se data seemed to suggest Miho styles (Class 2 decoration and pinched band at the<br />

neck were found mainly to the south <strong>of</strong> the wharf, and Gharanga/Kopo styles (punctate<br />

band) were found to the north <strong>of</strong> the wharf. A single sherd with a Honiavasa style Class<br />

1 bounded linear incised motif was found to the north <strong>of</strong> the wharf. <strong>The</strong> subsequent, more<br />

comprehensive, 1/5 sample transect surface collection enabled stylistic spatial structure to<br />

be established with more confidence. <strong>The</strong> distribution <strong>of</strong> ceramic styles at Zangana was<br />

compared to the overall quantity distribution. Wavy-lipped sherds and Gharanga/Kopo<br />

multi-band fingernail-pinched/punctate styles were distributed as expected from the<br />

distribution <strong>of</strong> the overall sherd sample (Figure 122, Figure 123).<br />

Deformation <strong>of</strong> the lip into a wave form<br />

(Sherd count)<br />

10<br />

5<br />

1<br />

New Georgia Mainland Roviana Lagoon<br />

Class 1 linear motif incised/fingernail-pinched sherds showed a different distribution to<br />

433<br />

35m<br />

�<br />

Figure 122: Lip deformation into a wave present in both<br />

Zangana-North and Zangana-South.


Band <strong>of</strong> punctation/multiple bands pinching<br />

(sherd count)<br />

10<br />

5<br />

1<br />

New Georgia Mainland Roviana Lagoon<br />

434<br />

35m<br />

�<br />

Figure 123: Bands <strong>of</strong> punctation were restricted to Zangana-North.<br />

the one expected from the overall sherd distribution, with sherds <strong>of</strong> this style being found<br />

mainly to the south <strong>of</strong> the wharf (Figure 124).<br />

<strong>The</strong> pattern expected from the overall distribution <strong>of</strong> all sherds would be for the<br />

bulk <strong>of</strong> Miho-style sherds to be found at the north end <strong>of</strong> the site, with only a few sherds<br />

found south <strong>of</strong> the wharf. Instead, 24 Miho sherds were found either near the wharf or on<br />

the reef flat to the south, with only four found at the north end if the site, where the<br />

majority <strong>of</strong> sherds were found. <strong>The</strong> chances <strong>of</strong> this happening randomly have not been<br />

calculated, but would be very slim.


Figure 124: Unbounded linear incision or necks banded with<br />

pinching at Zangana.<br />

Two alternate explanations for this patterning were considered, firstly that the Miho<br />

Class1/pinch style, the Miho “wav” lip deformation style, and the Gharanga/Kopo<br />

punctate/multi-band-pinch styles form a time-series, and are spatially segregated in this<br />

instance by a shift or expansion/contraction <strong>of</strong> the locus <strong>of</strong> settlement over time (possibly<br />

in the reverse order to that listed); secondly, that this patterning represents<br />

contemporaneous, spatially segregated stylistic diversity, interpretable as diversity <strong>of</strong><br />

manufacturing traditions or diversity <strong>of</strong> activity patterning. This second explanation seems<br />

unlikely, as the scale <strong>of</strong> the spatial strucure is too large to be activity patterning, and even<br />

if one end <strong>of</strong> a village were producing different pottery discard is unlikely to be so neatly<br />

structured.<br />

435


Zangana Temper Classes<br />

(sherd Count)<br />

70<br />

35<br />

7<br />

Quartz-calcite hybrid<br />

Placered volcanic<br />

Unplacered feldspathic<br />

New Georgia Mainland Roviana Lagoon<br />

436<br />

35m<br />

�<br />

Figure 125: Distribution <strong>of</strong> temper classes at Zangana.<br />

Distribution <strong>of</strong> Zangana Pottery Tempers:<br />

Temper class sherd counts did not show strong patterning across the site (Figure 125).<br />

Sherds tempered with unplacered volcanic/feldspathic/hornblendic stream sands are more<br />

common in the northern part <strong>of</strong> the site (see an enlarged detail <strong>of</strong> Figure 125 in Figure<br />

126). Quartz-calcite tempered sherds do not seem to have any particular departure from<br />

the pattern expected from overall sherd quantities, suggesting that this anomalous temper<br />

group is not highly spatially correlated with any <strong>of</strong> the decorative attributes. This suggests<br />

that, if the Zangana stylistic spatial structure represents a time series <strong>of</strong> styles, the quartz-<br />

calcite temper transcends this series, occurring throughout.


Zangana Temper Classes<br />

(sherd Count)<br />

100<br />

50<br />

10<br />

Unplacered feldspathic<br />

Placered volcanic<br />

Quartz-calcite hybrid<br />

Figure 126: Detail <strong>of</strong> distribution <strong>of</strong> temper classes at Zangana-<br />

North.<br />

Zangana Summary:<br />

In summary, at Zangana the pottery distribution shows evidence <strong>of</strong> wave size-sorting,<br />

where lag deposits <strong>of</strong> larger sherds are mainly found at the seaward margin <strong>of</strong> Zangana<br />

South. <strong>The</strong>re may be some settlement patterning in this part <strong>of</strong> the site, but the possibility<br />

<strong>of</strong> correlation with mictotopographic variation has yet to be examined, and random<br />

distribution might look something like the pattern observed. Stylistically, there is spatial<br />

structure at Zangana, with the Miho style having a different distribution to other styles.<br />

This may indicate a shifting locus <strong>of</strong> settlement across a ceramic time-series, or alternately<br />

a contemporaneous expression <strong>of</strong> manufacturing diversity across a settlement. None <strong>of</strong> the<br />

major temper types exhibit any clear spatial structure, although there is a (statistically<br />

untested) suggestion <strong>of</strong> a negative correlation between Miho style ceramics and unplacered<br />

volcanic temper.<br />

437


Hoghoi:<br />

Questions <strong>of</strong> taphonomy and other site formation questions at Hoghoi were addressed by<br />

mapping data from the second surface collection, which comprised a linear along-shore<br />

arrangement <strong>of</strong> 22 contiguous intertidal transects, each 5x12m, oriented with the long axis<br />

at right angles to the shoreline (the first surface collection the previous year had been a<br />

grab sample <strong>of</strong> 105 items, mostly decorated pottery).<br />

Natural Formation Processes at Hoghoi: Along-shore Size Sorting:<br />

<strong>The</strong> question: was there a smaller site in the past that has been spread into the present<br />

110m distribution by sherd/manuport transport? This was investigated by looking at the<br />

average size <strong>of</strong> sherds. Is there a spread <strong>of</strong> small sherds and a lag deposit <strong>of</strong> larger sherds?<br />

Or perhaps multiple spaced lag deposits <strong>of</strong> larger sherds interspersed with small sherds<br />

representing spread from multiple occupation loci? <strong>The</strong> overall sherd count per transect<br />

identifies those transects where sherd sample size is sufficient to address this question<br />

(Figure 127).<br />

Transect sample sizes<br />

(sherd count)<br />

150<br />

75<br />

15<br />

0 0 12 24 12 5 3 4 64 149 95 67 39 23 66 9 28 23 46 31 10 0<br />

0-5m 15-20m 30-35m 45-50m 60-65m 75-80m 90-95m 105-110<br />

Figure 127: Distribution <strong>of</strong> the sherd sample at Hoghoi.<br />

Transects 0 through 10m, 25 through 40m, 75-80m and 100 through 110m were excluded<br />

as having too small a sample <strong>of</strong> sherds to be reliable indicators <strong>of</strong> along-shore size-sorting<br />

processes.<br />

438<br />

�<br />

<strong>The</strong> fairly large sherd count at 70-75m is immediately adjacent to a transect with


a very low sherd count thought during field collection to have been the subject <strong>of</strong> coral<br />

boulder clearance for a copra-loading canoe-landing (the 70-75m transect appears to have<br />

been the recipient <strong>of</strong> the coral cleared from the 75-80m transect. It can be seen in figure<br />

12 that the average sherd area for these 66 sherds is noticeable smaller than for<br />

neighbouring transects, suggesting many <strong>of</strong> these sherds may have resulted from trampling<br />

and loss <strong>of</strong> wave-shelter in the “cleared” transect immediately to the west.<br />

Average sherd area<br />

(cm2, showing sample sizes)<br />

0-5m<br />

30<br />

15<br />

3<br />

0 12 24 12 5 3 4 64 149 95 67 39 23 66 9 28 23 46 31 10 0<br />

15-20m 30-35m 45-50m 60-65m 75-80m 90-95m 105-110<br />

Figure 128: Lack <strong>of</strong> sherd size variation at Hoghoi, except at 35-40m (n=4).<br />

439<br />

�<br />

<strong>The</strong>re is little evidence in Figure 128 that the site comprises concentrated lag-deposits <strong>of</strong><br />

large sherds flanked by transects containing only smaller sherds. <strong>The</strong> 45-50m sample <strong>of</strong><br />

149 sherds, which is the most dense scatter <strong>of</strong> sherds, is not composed <strong>of</strong> unusually large<br />

sherds, but the 50-60m transects have very slightly larger than average sherd area and<br />

reasonably high density <strong>of</strong> deposits.<br />

Lithics at Hoghoi:<br />

Data on lithic manuports, many <strong>of</strong> which appear to be ovenstones, was mapped to further<br />

investigate the evidence for along-shore size sorting. Lithic manuports tended to be larger<br />

on average between the 25m mark and 60m, in the region where sherd density is highest<br />

(Figure 129).


Average weight <strong>of</strong> Hoghoi Manuports (g)<br />

Sample count shown<br />

200<br />

100<br />

20<br />

3 3 3 11 0 8 20 9 44 53 11 29 21 12 14 8 30 12 19 23 7 4<br />

0-5m 20-25m 40-45m 60-65m 80-85m 100-105m<br />

Figure 129: Larger average manuport mass from 25m to 60m at Hoghoi.<br />

Of the smaller lithic manuports, many were unfractured water rounded cobbles, prompting<br />

speculation that these might have some non-cooking function such as net weights or sling-<br />

stones. Distribution by form and size classes is mapped in Figure 130.<br />

Hoghoi Manuport Forms/Sizes<br />

(Quantified by Count)<br />

40<br />

20<br />

4<br />

Rounded stones


showing little evidence <strong>of</strong> the strong clustering at 45-50m in evidence in the ceramic<br />

distribution, or the distribution <strong>of</strong> larger stones.<br />

It seemed reasonable to conclude that these smaller rounded stones were more<br />

transportable than ceramic sherds, and thus had been spread by wave action, obliterating<br />

any activity patterning. A lithic manuport type series, established through petrographic and<br />

macroscopic classification, identified several geological source groups within the<br />

ovenstone assemblage (Chapter 10). When manuports were mapped according to source<br />

group (Figure 131), Classes 3, 5 and 18 dominate at the eastern end <strong>of</strong> the site.<br />

Hoghoi Manuport Classes<br />

Quantified By Count<br />

30<br />

15<br />

Class 1<br />

Class 2<br />

Class 3<br />

Class 5<br />

Class 18<br />

3<br />

0-5m 20-25m 40-45m 60-65m 80-85m 100-105m<br />

Figure 131: Size-sorting <strong>of</strong> manuport petrographic classes at Hoghoi, suggestive<br />

<strong>of</strong> different size-procurement patterns by source.<br />

In addition to the size sorting effect there is also a possible difference in procurement<br />

pattern by source, with smaller stones being selected from some sources than others.<br />

Differences in the size <strong>of</strong> stones available or selected from different sources is likely to<br />

have produced the patterning in Figure 131. <strong>The</strong> only large Class 18 stone (1100g) was<br />

found in unit 10 (45-50m), <strong>of</strong>fering further support for a size-sorting wave-transport<br />

explanation for spatial structure at Hoghoi, although re-use <strong>of</strong> materials with remaining use<br />

life from an earlier, more easterly occupation cannot be ruled out.<br />

441


Distribution <strong>of</strong> Hoghoi Ceramic Tempers:<br />

Ceramic tempers <strong>of</strong>fer further evidence relating to the question <strong>of</strong> size/density sorting by<br />

waves as opposed to along-shore behavioural spatial structure (Figure 132). Quart-calcite<br />

and unplacered tempers are slightly more common from 65-80m than elsewhere, but since<br />

these tempers may have lower specific gravity than placered volcanic tempers with a high<br />

ferro-magnesian component, density sorting by along-shore wave processes could explain<br />

this patterning (but analysis showed no difference in specific gravity by temper).<br />

Distribution <strong>of</strong> pottery tempers<br />

(Quantified by sherd count)<br />

130<br />

65<br />

13<br />

Quartz-calcite hybrid temper<br />

Unplacered lithic-feldspathic<br />

Placered volcanic<br />

0-5m 20-25m 40-45m 60-65m 80-85m 100-105m<br />

Figure 132: Distribution <strong>of</strong> potsherd temper classes at Hoghoi.<br />

This analysis <strong>of</strong> the distribution <strong>of</strong> temper types suffers from sample size problems,<br />

and highlights the need to make the best possible use <strong>of</strong> the samples that have survived the<br />

last two-to-three millennia. <strong>The</strong> hundred or so larger sherds uplifted during the initial<br />

collection, together with the adzes collected at the same time, would have alleviated this<br />

problem to some extent.<br />

Distribution <strong>of</strong> Ceramic Styles at Hoghoi:<br />

<strong>The</strong> collection <strong>of</strong> a hundred decorated sherds or lithic artefacts from this site on first<br />

discovery has had a greater adverse effect on the stylistic component <strong>of</strong> this spatial analysis<br />

than on the manuport or temper analyses. This highlights the need for careful spatial<br />

recording on these fragile sites. <strong>The</strong>re is so little spatial stylistic information from the<br />

second collection at Hoghoi that the information was not mapped, and some general<br />

comments only are made.<br />

<strong>The</strong>re seems to be proportionately more deformation <strong>of</strong> the lip into a wave form<br />

442


and Class 2 linear motif on the rim or shoulder between 80m and 95m than between 40<br />

and 65m, especially at 80-85m,although numbers are so small that this could easily be<br />

sampling error or quantification error, but punctate band/multiple bands <strong>of</strong> pinching are<br />

found right along the site, albeit in small numbers ( peaking at six instances <strong>of</strong> punctate<br />

band at 45-50m).<br />

Hoghoi Summary:<br />

Sherd quantities suggest a strongly clustered distribution <strong>of</strong> potsherds in the vicinity <strong>of</strong> the<br />

45-50m transect, with a variable density across other transects. <strong>The</strong> largest mean sherd<br />

sizes, by a very small margin, were around the 45-50m transect, suggesting along shore<br />

size-sorting is operative to a slight extent, and that there is a slight lag effect in the vicinity<br />

<strong>of</strong> these larger (on average) sherds. <strong>The</strong> spatial distribution <strong>of</strong> lithic manuports also<br />

suggested along-shore size sorting (somewhat surprisingly in view <strong>of</strong> the higher specific<br />

gravity <strong>of</strong> the lithic manuports, but perhaps the flat shape <strong>of</strong> sherds makes them less subject<br />

to transport by rolling). Compositional differences and possible functional differences<br />

between manuport samples from the main 45-50m sherd cluster and the western sherd<br />

distribution are most likely a result <strong>of</strong> size-sorting by waves, added onto a difference by<br />

source in procurement strategies (small stones were taken from the Class 18 source for<br />

some reason).<br />

Initial collection <strong>of</strong> a hundred or so larger sherds from the site prior to more<br />

detailed transect collection made a spatial analysis <strong>of</strong> stylistic structure impossible, as these<br />

were the bulk <strong>of</strong> the sherds diagnostic <strong>of</strong> the styles defined in Chapter 9. This highlights<br />

the need to record the locations <strong>of</strong> items carefully for this fragile site type, which is easily<br />

rendered uninformative by casual surface collection. This finding is important for future<br />

resource management and research.<br />

Honiavasa:<br />

443


Sherds were collected in five pace-measured transects at Honiavasa, so the site maps on<br />

which the spatial analysis which follows are based are only roughly to scale, and transect<br />

areas may in practice have been unequal. Consequently, sherd counts cannot be translated<br />

into sherd densities, but relative frequencies <strong>of</strong> sherd attributes and taphonomic measures<br />

such as average sherd sizes can be used with confidence. Overall sherd sample sizes are<br />

reasonable for all five transects when looking at the distribution <strong>of</strong> sherd size and<br />

composition (Figure 133).<br />

Sherd Count<br />

110<br />

55<br />

11<br />

Honiavasa Passage<br />

� 35m<br />

444<br />

113<br />

98<br />

45<br />

76<br />

110<br />

Honiavasa Point<br />

Figure 133: Distribution <strong>of</strong> the sherd sample at Honiavasa.<br />

Taphonomy:<br />

Honiavasa has the largest mean sherd size <strong>of</strong> all sites in the study. <strong>The</strong> site had not<br />

previously been discovered by archaeologists. Honiavasa is located opposite the modern<br />

village <strong>of</strong> Sasavele, across the reef passage, and is not subject to intensive foot traffic or<br />

intensive children’s play activities. <strong>The</strong> site as collected was largely subtidal in water up<br />

to a metre deep at low water on 20 September 1997, and may extend into deeper water.<br />

This means that for most <strong>of</strong> the year the site is well submerged and sherds are seldom<br />

subject to high-energy wave transport, despite the reef-passage location. Sherd size<br />

increases slightly with increasing depth <strong>of</strong> water from east to west (Figure 134). <strong>The</strong><br />

centre transect is that most exposed to wave -induced sherd damage, as it is here the reef<br />

flat proper begins, and here that refracted swell curling around from the passage to the east<br />

breaks at low tide. This is reflected in sherd size, although low collection intensity in this<br />

transect due to a rising tide may have been a factor.


Average sherd area (cm2)<br />

33<br />

16.5<br />

3.3<br />

Honiavasa Passage<br />

� 35m<br />

445<br />

Honiavasa Point<br />

Figure 134: Effects <strong>of</strong> wave refraction (and collection intensity?) on sherd<br />

size at Honiavasa: the western margin is exposed to waves from Honiavasa<br />

channel, which expend their energy in a swash zone at about the centre <strong>of</strong><br />

the site at low tide.<br />

Distribution <strong>of</strong> Temper Types:<br />

Transect 2 (2 nd from right, n = 76) shows an increased frequency <strong>of</strong> placered volcanic<br />

/unplacered pyroxenic tempers (Figure 135), otherwise the relative proportions <strong>of</strong> the<br />

three major temper classes are reasonably consistent across the site. (Placered volcanic<br />

with volcanic rock fragments is a subgroup <strong>of</strong> the placered volcanic group, and difficult<br />

to distinguish reliably without petrographic analysis. Variations in the relative frequencies<br />

in which this temper was identified may have more to do with variations in surface etching<br />

resulting from cleaning than with behaviour in the past, hence shown in the same shade as<br />

placered volcanic temper.)<br />

Temper Types<br />

110<br />

55<br />

11<br />

Placered volcanic temper<br />

Pv with V.rock fragments<br />

Unplacered lithic/feldspathic<br />

Quartz-calcite<br />

Roviana Lagoon<br />

�<br />

Honiavasa Point<br />

0 30m<br />

Figure 135: Distribution <strong>of</strong> pottery tempers at Honiavasa.


Figure 136: Co-joining sherds from deeper western margin <strong>of</strong> Honiavasa..<br />

<strong>The</strong>re is a pattern <strong>of</strong> slightly increased incidence <strong>of</strong> quartz-calcite tempers at the eastern<br />

and western margins <strong>of</strong> the site. Analysis <strong>of</strong> vessel completeness did show, however, that<br />

the western margins <strong>of</strong> the site yielded larger sherds, and one particularly complete vessel<br />

including the only two co-joining sherds recovered during the entire study (Figure 136),<br />

so perhaps taphonomic variation should not be hastily discounted in explaining variation<br />

in the distribution <strong>of</strong> Quartz-calcite tempered sherds. <strong>The</strong> same could hold for unplacered<br />

lithic/feldspathic sherds, which might be slightly weaker, hence smaller and more easily<br />

transported than placered volcanic sherds.<br />

Distribution <strong>of</strong> Decorative Attributes at Honiavasa:<br />

Due to the relative plain-ness <strong>of</strong> the Honiavasa pottery, analysis <strong>of</strong> the spatial distribution<br />

<strong>of</strong> styles at Honiavasa is faced with severe sample size problems (Figure 137). <strong>The</strong><br />

relatively high frequency <strong>of</strong> notched lips <strong>of</strong> two types, and relatively low frequency <strong>of</strong><br />

carinations/Class 1 linear motifs at the western end <strong>of</strong> the site may be a sampling error,<br />

446


Distribution <strong>of</strong> decorative attributes<br />

(Quantified by sherd count)<br />

30<br />

15 3<br />

Band <strong>of</strong> impression on the top face <strong>of</strong> the lip, or on the outer edge<br />

Carinated vessels or Class 1 linear motifs<br />

Band <strong>of</strong> fingernail pinching on the outer edge <strong>of</strong> the lip<br />

Roviana Lagoon<br />

�<br />

Honiavasa Point<br />

0 30m<br />

Figure 137: Distribution <strong>of</strong> pottery decorative attributes at<br />

Honiavasa.<br />

especially in view <strong>of</strong> higher vessel completeness in transect 5.<br />

Honiavasa Summary:<br />

While sherd size distribution suggests some taphonomic variation across the site consistent<br />

with field observation <strong>of</strong> site morphology and wave exposure, the data on Honiavasa<br />

temper distribution suggests that mean sherd size as a taphonomic indicator may mask<br />

additional significant taphonomic variation. Recovery <strong>of</strong> two large cojoining sherds from<br />

transect five at the deeper western margin <strong>of</strong> the site alongside Honiavasa passage provides<br />

a clue that taphonomic degradation has been less severe in this part <strong>of</strong> the site. If<br />

preservation is better at transect 5 this may indicate that the relatively large sample <strong>of</strong><br />

sherds from that deeper location actually represents a relatively small number <strong>of</strong> vessels<br />

in a more complete state, a biased sample from which several diagnostic pieces may derive<br />

from a single vessel.<br />

Chapter Summary and Conclusions.<br />

447


In drawing together the information from analyses <strong>of</strong> these three sites, the first issue is<br />

whether an inter-site comparison <strong>of</strong> any <strong>of</strong> the themes <strong>of</strong> these analyses, size distribution,<br />

temper distribution, or stylistic distribution, can be informative. A question in this regard<br />

mentioned above in relation to Zangana stylistic distribution was whether a comparison <strong>of</strong><br />

the distribution <strong>of</strong> styles between sites would show up regularities in such patterning, or<br />

whether these might be considered as random variation from site to site. A further question<br />

for intersite comparisons that arose in relation to a structured distribution <strong>of</strong> tempers at<br />

Hoghoi was whether the anomalous quartz-calcite temper group separated out spatially<br />

at all in other sites. <strong>The</strong> Honiavasa site analysis raised questions about the methods used<br />

to assess intrasite taphonomic variation, the significance <strong>of</strong> which can now be discussed<br />

in relation to other sites.<br />

Class 2 Linear Motifs and Neck Pinching:<br />

<strong>The</strong>re was a clear pattern <strong>of</strong> spatial segregation <strong>of</strong> Class 2 linear motifs or neck pinching<br />

(Miho style) from punctate bands/multi-band pinching (Gharanga/Kopo style) at Zangana,<br />

which justifies Zangana South being treated separately to Zangana North in seriation<br />

(Chapter 12). At Hoghoi, no firm conclusion <strong>of</strong> this nature could be drawn, and these<br />

styles do not occur at Honiavasa. Of the two other sites where the Miho style occurs there<br />

is no repetition <strong>of</strong> the pattern observed at Zangana, so the hypothesis that this spatial<br />

structure represents a contemporaneous socio-ceramic spatial structure is rejected. This<br />

leaves the competing alternative hypothesis, that the Miho style and the Gharanga/Kopo<br />

styles form a time-series, and are spatially segregated in this instance by a shift or<br />

expansion/contraction in the locus <strong>of</strong> settlement over time. This conclusion is significant<br />

for seriations in Chapter 12.<br />

Spatial Separation <strong>of</strong> Quartz-Calcite Temper:<br />

At Hoghoi, the distribution <strong>of</strong> quartz-calcite temper showed some structure, but this did<br />

not seem to be the case at Zangana. At Honiavasa there was a hint <strong>of</strong> similar structure, and<br />

448


in neither case Honiavasa or Hoghoi) can taphonomic sorting or simply small-sample error<br />

be ruled out as an explanation.<br />

Implications for Resource Management and Future Research:<br />

Spatial structure in collection sites has been shown to be key to interpreting these sites.<br />

Wave sorting processes and swash-zone taphonomic effects can be seen in all three sites<br />

analysed in this <strong>chapter</strong>. Also, the difficulty in analysing stylistic spatial structure at Hoghoi<br />

was increased by removal <strong>of</strong> an initial grab sample. Similar sampling strategies at Paniavile<br />

by a series <strong>of</strong> archaeological teams and by locals meant that no spatial analysis could be<br />

attempted there.<br />

It is clear that future research and resource management <strong>of</strong> such surface sites can<br />

usefully apply detailed spatial recording, to the level <strong>of</strong> point-provenancing <strong>of</strong> artefacts and<br />

recording <strong>of</strong> micro-topography, otherwise formation processes <strong>of</strong> these scatters <strong>of</strong> pottery<br />

are unlikely to be understood. It seems unlikely to me that some <strong>of</strong> the fundamental<br />

difficulties in seriating surface collections, such as unraveling temporal mixing from<br />

different periods, and distinguishing activity patterning from taphonomic spatial sorting or<br />

microtopographic trapping, can be resolved satisfactorarily unless such recording is done<br />

carefully.<br />

449


450


Introduction:<br />

CHAPTER 12:<br />

CHRONOLOGY<br />

In this <strong>chapter</strong> information from thermoluminescence, radiocarbon determinations, and<br />

ceramic seriation is presented, followed by a discussion which attempts to draw this<br />

diverse information into a provisional chronology. Implications for future research and<br />

resource management are also discussed.<br />

Seriation measures time as a continuum <strong>of</strong> gradual change, in that rates <strong>of</strong> change<br />

cannot be measured by seriation. In seriation chronological types or attributes are treated<br />

as arbitrary units <strong>of</strong> measurement <strong>of</strong> the process <strong>of</strong> change rather than as essential classes<br />

in the data, allowing change over time between one type and another (O'Brien & Lyman<br />

2000:52-56). <strong>The</strong> seriator is not interested in establishing fixed boundaries <strong>of</strong> a type, for<br />

example, whether something is Lapita or not, but rather in discovering which aspects <strong>of</strong><br />

artefact variability can be used to construct a coherent evolutionary series with which to<br />

tell time. By contrast, in using absolute dates one is usually (but not always) concerned to<br />

know which chunk <strong>of</strong> time is being dated (which event). Paradoxically, although<br />

radiocarbon age is a continuous variable, applications <strong>of</strong> the technique seek to date<br />

occupations, layers, horizons, and thus seek a discontinuous measure <strong>of</strong> time (for an<br />

extended discussion <strong>of</strong> this see O'Brien & Lyman 2000). “Is this Lapita? How old is type<br />

X? When was this stratigraphic unit deposited? When does Lapita end? Absolute dating<br />

and seriation thus work best when combined in a dialogue in which the flow <strong>of</strong> time within<br />

and between events can be calibrated by the age <strong>of</strong> the events and inform on the properties<br />

<strong>of</strong> the events, such as relative order.<br />

451


In this <strong>chapter</strong> the radiocarbon dates and thermoluminescence dates used are direct<br />

dates associated with the manufacture or use <strong>of</strong> particular pots. Sites are temporal<br />

mixtures, but the direct dates give the age <strong>of</strong> pots, not sites, and thus partially escape the<br />

discontinuous view <strong>of</strong> time <strong>of</strong> dating by stratigraphic association.<br />

Patchy sampling <strong>of</strong> a ceramic series might look like natural groupings in the<br />

ceramic data, which could be misinterpreted as culture-historical phases. Distinguishing<br />

between these alternatives, patchy sampling and discrete phases <strong>of</strong> occupation, requires<br />

that other lines <strong>of</strong> evidence besides ceramic variability be taken into account to assess<br />

whether there is historical continuity and heritable continuity in the sample.<br />

If, for example, at some period the action moved <strong>of</strong>fstage, out <strong>of</strong> the sampling<br />

frame, and then descendants <strong>of</strong> the same people returned later, while there might be a<br />

discontinuity in the sampled ceramic evolutionary record, other factors such as type <strong>of</strong><br />

settlement or use <strong>of</strong> raw materials and adze maufacturing techniques might be more<br />

conservative, and provide evidence for heritable continuity.<br />

<strong>The</strong> Roviana ceramic data falls quite easily into styles or types (as discussed in<br />

Chapter 9). This tendency <strong>of</strong> decorative variability to come in types is regarded below as<br />

more likely to be telling us something about the patchy view <strong>of</strong> time presented by the<br />

available sample, than indicating any sort <strong>of</strong> cultural replacement. <strong>The</strong> argument is made<br />

below that there is evidence for heritable continuity across the entire sample, from<br />

settlement location type and from adzes, as well as from occasional sherds that are<br />

intermediate between the styles or types that most decorated sherds can be easily assigned<br />

to. <strong>The</strong> tendency for sherds to group neatly most likely results from historical<br />

discontinuities in the sample rather than any lack <strong>of</strong> heritable continuity. This does not<br />

automatically mean that there were hiatuses in occupation, just that the sample available<br />

to the analyst is historically incomplete, in that some periods are not well represented.<br />

452


Possibly settlements from these periods have been destroyed by waves due to exposed<br />

location, are hidden under s<strong>of</strong>t sediments, or simply have yet to be found, or alternatively<br />

people may have indeed have been largely absent from the sampling frame at some times.<br />

We cannot discriminate between these possibilities on current evidence.<br />

Radiocarbon data from a series <strong>of</strong> AMS dates is presented and discussed in the<br />

following section, followed in turn by a section on TL dating, and a section on seriation.<br />

TL dates do not give any absolute chronology, but provide useful data on the relative ages<br />

<strong>of</strong> units used in the seriation, despite problems with anomalous fading <strong>of</strong> the TL signal<br />

from the feldspar grains used in the analysis. A series <strong>of</strong> CA seriations are performed,<br />

examining the effect <strong>of</strong> controlling for vessel form in the selection <strong>of</strong> attributes used.<br />

AMS Radiocarbon Dates on Potsherds:<br />

As pottery was collected from the lower margins <strong>of</strong> the swash zone predominantly, spatial<br />

associations between organic materials and the pottery were unlikely to be temporal<br />

associations. Four AMS radiocarbon dates were obtained from sherds (Table 46). <strong>The</strong>se<br />

involve three different modes <strong>of</strong> association between pot and dated carbon,<br />

• the formation <strong>of</strong> the clay deposit from which the pot was made,<br />

• mixing <strong>of</strong> the clay with environmental impurities during manufacture <strong>of</strong> the vessel<br />

• accumulation <strong>of</strong> smoke-derived carbon on the surface <strong>of</strong> the vessel during use.<br />

Also, there is evidence from two <strong>of</strong> the dates to suggest that distinguishing between all <strong>of</strong><br />

these modes <strong>of</strong> association is possible, suggesting direct dating <strong>of</strong> pottery styles is now a<br />

practical reality in Oceanic archaeology.<br />

453


Table 46: Radiocarbon determinations from pottery (calibrated using OxCal 3.5,<br />

Stuiver et al. 1998 atmospheric data).<br />

Sample ID Fraction<br />

Dated<br />

Lab<br />

Number<br />

HV.4.143 “seed” NZA-<br />

12345<br />

HV.4.143 blackened<br />

edge<br />

P.68 charcoal<br />

inclusion<br />

HG.15s.1 fine soot on<br />

sherd<br />

surface<br />

NZA-<br />

12346<br />

Years BP Cal. Age<br />

2sigma<br />

8250+/-370 8300BC-<br />

6400BC<br />

9742+/-80 9350BC-<br />

8800BC<br />

AA33504 2130+/-90 390BC-<br />

30AD<br />

NZA-1253 2619=+/-<br />

45<br />

454<br />

900BC-<br />

550BC<br />

Site Unit<br />

Honiavasa 4<br />

Honiavasa 4<br />

Paniavile Jan 1996<br />

Hoghoi 15s<br />

<strong>The</strong> dates from sherd HV.4.143 (Figure 138) most likely date the geomorphological<br />

formation <strong>of</strong> the clay deposit from which the pottery was eventually made. <strong>The</strong> round<br />

“seed” (identity as a seed is unconfirmed) dated to within 500 years in age <strong>of</strong> a dark<br />

carbon-rich zone near the surface <strong>of</strong> the sherd, visible in the break. <strong>The</strong> carboniferous zone<br />

can be expected to contain small amounts carbon expelled from the clay on firing,<br />

potentially a mix <strong>of</strong> old clay-derived carbon and new carbon mixed in at manufacture,<br />

while the round “seed” could have been included either at the time <strong>of</strong> manufacture <strong>of</strong> the<br />

pot or at the time <strong>of</strong> geological formation <strong>of</strong> the clay. As both ages are substantially older<br />

than the expected age <strong>of</strong> the pottery from style and provenance, we can be reasonably<br />

confident that these do not date the manufacture <strong>of</strong> the pot.<br />

<strong>The</strong> charcoal inclusion in sherd P.68 from Paniavile must predate firing <strong>of</strong> the<br />

vessel and is <strong>of</strong> an age broadly consistent with our expectations <strong>of</strong> the approximate<br />

antiquity <strong>of</strong> the intertidal pottery series (circa-or-post-Lapita) (Figure 139).


Figure 138: A method <strong>of</strong> identifying sub-fossil organic inclusions?<br />

(Image supplied by Rafter Radiocarbon Lab)<br />

Radiocarbon determination<br />

2600BP<br />

2400BP<br />

2200BP<br />

2000BP<br />

1800BP<br />

1600BP<br />

Atmospheric data from Stuiver et al. (1998); OxCal v3.5 Bronk Ramsey (2000); cub r:4 sd:12 prob usp[chron]<br />

AA33504 : 2130±90BP<br />

68.2% probability<br />

360BC (15.3%) 290BC<br />

240BC (52.9%) 40BC<br />

95.4% probability<br />

390BC (95.4%) 30AD<br />

1000CalBC 500CalBC CalBC/CalAD<br />

Calibrated date<br />

500CalAD<br />

Figure 139: Calibration <strong>of</strong> radiocarbon determination from a charcoal inclusion in<br />

a sherd from Paniavile.<br />

455


Sherd HG.15s.1 was dated using smoke-derived carbon which had formed a fine soot on<br />

the surface <strong>of</strong> the sherd, preserved as a result <strong>of</strong> shallow burial in s<strong>of</strong>t mud (Figure 140,<br />

Figure 141, Figure 142). <strong>The</strong> sherd had been treated by the archaeologist with methyl<br />

methacrylate (B72). Pretreatment at the Rafter lab included soaking in acetone, scraping<br />

<strong>of</strong>f <strong>of</strong> a soot sample, followed by a series <strong>of</strong> organic solvent washes (two cycles <strong>of</strong> hexane,<br />

two cycles <strong>of</strong> isopropanol and four cycles <strong>of</strong> acetone, prior to the usual acid/alkali/acid<br />

sequence) (Prior 2000). This allows a reasonable level <strong>of</strong> confidence that the methyl<br />

methacrylate was removed and the 14C age is reliable. <strong>The</strong> extent to which the soot dates<br />

the use <strong>of</strong> the pot is dependent on the average age <strong>of</strong> the fuel used in the fires, so there is<br />

some potential for an old wood effect.<br />

Radiocarbon determination<br />

2900BP<br />

2800BP<br />

2700BP<br />

2600BP<br />

2500BP<br />

2400BP<br />

2300BP<br />

2200BP<br />

Atmospheric data from Stuiver et al. (1998); OxCal v3.5 Bronk Ramsey (2000); cub r:4 sd:12 prob usp[chron]<br />

NZA12353 : 2619±45BP<br />

68.2% probability<br />

830BC (65.2%) 780BC<br />

775BC ( 3.0%) 765BC<br />

95.4% probability<br />

900BC (86.7%) 750BC<br />

690BC ( 3.5%) 660BC<br />

640BC ( 3.6%) 590BC<br />

580BC ( 1.6%) 550BC<br />

1000CalBC 800CalBC 600CalBC 400CalBC<br />

Calibrated date<br />

Figure 140: Calibration <strong>of</strong> a radiocarbon determination from smoke-derived<br />

carbon on a sherd from Hoghoi.<br />

<strong>The</strong> dates from Paniavile and Hoghoi do not overlap at 95% confidence interval, and<br />

suggest that the total duration <strong>of</strong> the intertidal/stilt house settlement pattern was at least<br />

160 years and possibly much longer. <strong>The</strong>se dates suggest there is likely to be significant<br />

456


temporal variability within the total intertidal-zone pottery sample, which is good news for<br />

seriation, as the possibility that the decorative variation which forms the data <strong>of</strong> the<br />

seriation is contemporaneous non-temporal variability <strong>of</strong> some sort is reduced.<br />

<strong>The</strong> older <strong>of</strong> these two dates is from Hoghoi, while the younger is from Paniavile,<br />

but as each <strong>of</strong> these sites contains a mixture <strong>of</strong> styles, and as neither <strong>of</strong> the dated sherds<br />

is clearly identifiable to particular styles, these dates don’t tell us much about the relative<br />

ages <strong>of</strong> styles. Although the Hoghoi site has a lot <strong>of</strong> Gharanga-Kopo style sherds, sherds<br />

<strong>of</strong> the dated sample are from an area <strong>of</strong> the site (transects 12 and 15) where attributes<br />

identified with the Miho style occur also, and the form <strong>of</strong> the vessel, with tallish thin<br />

Figure 141: AMS sample taken from blackened sherd at far left, found on<br />

the surface <strong>of</strong> unit 12 at Hoghoi. <strong>The</strong> sherd to the right, found in a<br />

subsurface test <strong>of</strong> unit 15 at Hoghoi, may be from the same vessel, and also<br />

has a sooted surface.<br />

excurvate rim, is more like the Miho style than Kopo style (Figure 141, Figure 142). <strong>The</strong><br />

only decoration, if it can be called that, is a deep impressed groove around the neck. <strong>The</strong><br />

vessel is unusually small in horizontal dimensions, which may have been a factor<br />

contributing to its preservation through increased form strength.<br />

457


Figure 142: Vessel with surface sooting from Hoghoi dated by AMS<br />

radiocarbon.<br />

Another sherd in the illustrations thought to be from this vessel was located by test<br />

excavation within 20m <strong>of</strong> the found location, on the surface in square 12. <strong>The</strong>se were<br />

attributed to the same vessel by virtue <strong>of</strong> similar form, the grooving at the neck, and similar<br />

sooting on the body surface, not seen on any other sherds.<br />

<strong>The</strong> dated sample from Paniavile was recovered from a plain sherd comprising a<br />

s<strong>of</strong>t shoulder and globular body, which cannot be identified to any stylistic grouping or<br />

period.<br />

<strong>The</strong>rmoluminescence (TL) Dates from Quartz-Calcite Sherds:<br />

Applicability <strong>of</strong> TL:<br />

<strong>The</strong>rmoluminescence dates the time a pot was last heated to a sufficient temperature to re-<br />

set the luminescence clock (about 450/C) (Feathers 1997). In most cases this will date<br />

either the manufacture <strong>of</strong> the pot or its last use. This assumption is particularly secure for<br />

materials discarded into the sea. This makes TL particularly useful for surface<br />

archaeological distributions in the intertidal zone and in shallow water.<br />

458


Uncertainties arising from factors affecting the precision <strong>of</strong> radiocarbon dates such as old<br />

carbon and contamination from geologic carbon do not apply to thermoluminescence.<br />

Calibration <strong>of</strong> radiocarbon dates sometimes involves a reduction in precision, when the<br />

calibration curve is flat, as in the period 760BC-250BC (Figure 138 and Figure 140). This<br />

flatness makes charcoal-dating the transformations <strong>of</strong> the Lapita cultural complex<br />

following its initial spread across Oceania extremely difficult.<br />

<strong>The</strong> disadvantage <strong>of</strong> thermoluminescence dating lies in “its dependence on<br />

numerous complex variables that can be difficult to estimate in any given situation”<br />

(Feathers 1997:4). TL dating can be used as an ordinal measure <strong>of</strong> age, an interval measure<br />

or as a ratio scale measurement, with a reduction in precision as one moves up the scale<br />

between these. Long-duration events such as an occupation can be dated with more<br />

precision by dating more <strong>of</strong> the objects involved (Feathers 1997:7).<br />

Luminescence Physics:<br />

<strong>The</strong> following is a synthesis from a recent TL review (Feathers 1997:11-54) with<br />

comments relevant to the Roviana case added.<br />

A material that has been exposed to natural radiation will emit a faint light when<br />

heated, proportional to the amount <strong>of</strong> energy absorbed, and proportional to time elapsed<br />

since such accumulation began. This relationship breaks down when enough time has<br />

elapsed for the material dated (the dosimeter) to become saturated with radiation, to the<br />

point where it can trap no more energy. <strong>The</strong> most common natural radioactive isotopes are<br />

40 K, 238 U and 232 Th, which have long half-lives and effectively constant global levels,<br />

allowing estimation <strong>of</strong> dose rates impinging on archaeological samples. <strong>The</strong>re are other<br />

minor contributors to natural radiation flux. Terrestrial natural radiation can be divided into<br />

three kinds (Alpha, Beta, and Gamma rays) <strong>of</strong> which Beta and Gamma rays are effective<br />

in inducing luminescence, by virtue <strong>of</strong> their penetrating power (Alpha rays do affect the<br />

surfaces <strong>of</strong> grains and small grains).<br />

Successful dating requires a source <strong>of</strong> radiation in the environment <strong>of</strong> the item to<br />

be dated, or even within the item itself, and a natural dosimeter within the item (which<br />

459


may also be a source <strong>of</strong> radiation, as in the cases <strong>of</strong> potassium feldspar, calcite or zircon).<br />

Quartz and sodic feldspars are good dosimeters, but lack an internal radiation source.<br />

None <strong>of</strong> the usual dosimeters will become saturated in less than about 100Ka (thousand<br />

years) so saturation is not a serious problem for Lapita archaeology.<br />

Fading is said to occur when the accumulated latent signal is not stable through<br />

time, but leaks away at ambient temperatures, making the date too young, either by<br />

mechanisms for which the probability <strong>of</strong> fading is known (thermal fading) or by poorly-<br />

understood and difficult-to-predict mechanisms (anomalous fading). <strong>The</strong>rmal fading can<br />

be more <strong>of</strong> a problem in hot climates, and for older samples, but in the case <strong>of</strong> the Roviana<br />

pottery thermal fading is less <strong>of</strong> a factor as the pottery has been underwater for most <strong>of</strong> its<br />

relatively short depositional history and has not been subject to high ambient temperatures.<br />

Anomalous fading can render a sample undatable, and is thought to have affected the<br />

Roviana samples, being a major obstacle to accurate calendrical dating in this case.<br />

Anomalous fading is more common with feldspars, and seldom affects quartz, unless the<br />

latter has feldspar impurities.<br />

To derive the age <strong>of</strong> the sample two measurements are required, equivalent dose<br />

and annual dose. <strong>The</strong> age is the equivalent dose (the total amount <strong>of</strong> radiation absorbed<br />

since last zeroed) divided by the annual dose. Equivalent dose is measured either by the<br />

additive dose technique, whereby the material is irradiated in increments and the growth<br />

curve <strong>of</strong> the emitted luminescence is compared to that <strong>of</strong> a standard, or by the regeneration<br />

technique, whereby the luminescence <strong>of</strong> the sample is measured, then zeroed by heat or<br />

light, then rebuilt by incremental dosing, producing a regeneration curve. <strong>The</strong> shapes <strong>of</strong><br />

these response curves (the sensitivity <strong>of</strong> the sample) are used in both techniques to<br />

estimate the age <strong>of</strong> the sample. Sensitivity to dose can change during measurement.<br />

Linearity in the growth curve cannot be assumed. <strong>The</strong> causes <strong>of</strong> sensitivity changes during<br />

measurement are not all well understood. After a zeroing event, the luminescence signal<br />

can build at a different rate, depending on the nature <strong>of</strong> the zeroing event.<br />

Measuring emission is more difficult for young samples like pottery, as the<br />

amount <strong>of</strong> light emitted is less. Selection <strong>of</strong> an appropriate spectral filter may be required<br />

460


to maximise the signal from such samples. Detection equipment tends to be designed for<br />

older samples than typically analysed by archaeologists, and may lack sufficient resolution<br />

for the task in some labs.<br />

Measuring dose rate requires correction for moisture content, correction for<br />

uncertainty regarding the external dose rate, and the possibility <strong>of</strong> changes in the dose rate<br />

through time. Water absorbs some <strong>of</strong> the radiation, and correction for this factor is easiest<br />

where the environment is known to have been either saturated or fully dry during the time<br />

since zeroing (most likely fully saturated in the Roviana case). Alpha and beta radiation,<br />

with little penetrating ability, are relatively easy to estimate from the properties <strong>of</strong> the<br />

sample itself, but gamma radiation, with a range <strong>of</strong> up to 30cm in soil, comes mostly from<br />

outside the sample, requiring some idea <strong>of</strong> the homogeneity <strong>of</strong> the environment in this<br />

respect. <strong>The</strong> Roviana sites with a scatter <strong>of</strong> volcanic ovenstones and volcanic-tempered<br />

pottery in a predominantly calcareous environment, are potentially inhomogeneous in this<br />

respect.<br />

“Particularly troublesome is the presence within the 30cm sphere <strong>of</strong> large<br />

clasts with a radioactivity different from that <strong>of</strong> the surrounding sediment”<br />

(Feathers 1997:31).<br />

Placement <strong>of</strong> zeroed dosimeters in the sample location, or use <strong>of</strong> portable gamma-<br />

ray scintillometers can calculate the modern dose to which the sample is exposed, but<br />

where there is a history <strong>of</strong> turbation or movement <strong>of</strong> the sample, calculation <strong>of</strong><br />

environmental gamma dose will not be as clear-cut, but may be less variable between<br />

samples from a site due to an averaging effect over time. Perhaps the issues for the<br />

Roviana materials is how long have these been in lag deposits as opposed to a stable<br />

matrix, and whether the dose rate was lower or more variable between sherds when these<br />

were deeply immersed, compared to the present dose rate. A detailed sea-level history<br />

combined with placement <strong>of</strong> dosimeters in experimental deposits could partially answer<br />

these questions.<br />

Weathering can affect internal and external dose rate through postdepositional<br />

movement <strong>of</strong> radionuclides. For example water can remove U but leave Th intact. Clay<br />

461


minerals in ceramics are end products <strong>of</strong> weathering, and in the case <strong>of</strong> the Roviana<br />

ceramics, have clearly been exposed to salt water. Low fired ceramics (less than 1000/C,<br />

as in the Roviana case as calcite reef detritus is unsintered), are particularly prone to these<br />

effects, but if such a disequilibrium is detected for radionuclides with short half-lives a<br />

correction can sometimes be applied, especially where the leaching effect is more or less<br />

constant. If radionuclides <strong>of</strong> longer half-life are involved this is more difficult, and<br />

uncertainty about dose rate can be the result. Detection <strong>of</strong> such disequilibrium is largely<br />

a matter <strong>of</strong> cost: a variety <strong>of</strong> measurement techniques can be applied, the cheaper <strong>of</strong> which<br />

measure alpha radiation only, and estimate other radiation from elemental abundance<br />

combined with natural isotopic abundance, or by Neutron Activation Analysis. High-<br />

resolution gamma spectroscopy is expensive, requires long counting times and large<br />

samples, but measures the gamma activities <strong>of</strong> several radionuclides, and this supplies a<br />

means to assess equilibrium.<br />

Luminescence signal can be measured from individual grains (inclusion dating),<br />

usually from either quartz or feldspar, or, where possible, zircon, which has abundant<br />

radioactive impurities that negate the need for estimation <strong>of</strong> external dose. Use <strong>of</strong><br />

individual grains facilitates choice <strong>of</strong> dating procedures and eliminates influence from clay<br />

minerals. Fine grain analysis, making no attempt to separate particular minerals, is required<br />

where large grains are unavailable, and despite some problems, can be useful where the<br />

environmental dose is complicated because the influence <strong>of</strong> gamma radiation is less with<br />

the inclusion <strong>of</strong> alpha-irradiated fine grains.<br />

Low signal output is the commonest cause <strong>of</strong> low precision when dating young<br />

ceramics. <strong>The</strong> best results come from samples in which there are abundant quartz grains<br />

<strong>of</strong> optimum size. Rocks which have been sufficiently heated <strong>of</strong>ten produce reliable dates,<br />

using either TL or OSL. <strong>The</strong> main reason lithic materials are seldom used for TL dating<br />

is that there is seldom reason to believe they have been sufficiently heated, although there<br />

are also greater sample preparation problems with lithics. I see no reason why ovenstones<br />

should not have been heated as highly if not higher than open-fired calcite-tempered<br />

pottery.<br />

462


Roviana TL Data:<br />

Eight pottery samples from four archaeological sites were submitted to the <strong>University</strong> <strong>of</strong><br />

Washington Anthropology Department TL laboratory. All were from the quartz-calcite<br />

temper group, which held the best prospect for isolating grains suitable for TL dating<br />

(Feathers 2002).<br />

Environmental Dose:<br />

Nine sediment samples and one ovenstone sample were measured to aid in the estimation<br />

<strong>of</strong> environmental radiation. Radioactivity <strong>of</strong> sherds was higher than that <strong>of</strong> sediments<br />

samples. Volcanic manuports from Zangana yielded higher radioactivity measurements<br />

than the calcite reefal detritus samples. Environmental dose rate was calculated by<br />

averaging the dose from the sediment samples for each site, with the manuport<br />

radioactivity omitted.<br />

Equivalent Dose and Fading:<br />

Two <strong>of</strong> the smaller samples, one from Gharanga and one from Miho, which also lacked<br />

sufficient quartz gains for inclusion dating, were analysed by the fine-grain method, and<br />

yielded equivalent dose measurements <strong>of</strong> low precision, due to weak signal-to-noise ratio.<br />

<strong>The</strong> other sherds were initially analyzed using the inclusion method on extracted quartz,<br />

but unreasonably old ages with high error terms were obtained, leading Feathers to<br />

conclude there was a spurious luminescence component, and this approach was aborted.<br />

Inclusions obtained by density separation <strong>of</strong> Potassium feldspars and Aluminium<br />

Silicates produced a sufficiently strong signal for dating, but four <strong>of</strong> theses samples showed<br />

evidence for anomalous fading (which can make the age too young). A correction for<br />

fading was applied to three <strong>of</strong> these samples, which resulted in low precision. Results are<br />

given in Table 47.<br />

Feathers rejected the fading correction for uw476 as it seemed too old based on<br />

the information supplied to him, which was that the occupation being dated ran from<br />

463


about 2500-1900BP, however, the corrected age, as initially supplied by him by email, is<br />

entirely consistent with the expected age <strong>of</strong> the Honiavasa Lapita site, and is therefore<br />

included in Table 47. Looking at the ages uncorrected for fading, the two dates from<br />

Honiavasa were the oldest (uw475 and uw476), while two <strong>of</strong> the dates from Paniavile<br />

(UW 472 and uw474) and the date from Gharanga (uw471) formed a young group. <strong>The</strong><br />

other date from Paniavile could be grouped with the Miho date and the Zangana south date<br />

( uw473, uw477, uw478). While Feathers does not think the dates can be used as relative<br />

ages as they are statistically indistinguishable when corrected for fading, the argument will<br />

be made below that the uncorrected ordination <strong>of</strong> ages, ignoring their standard deviations,<br />

agrees with the seriation favoured below, and can be reconciled with the AMS C14<br />

evidence also. This correlation with the seriation additionally supports the inference drawn<br />

from the decorative analysis <strong>of</strong> the quartz-calcite hybrid temper, which was that this temper<br />

group was imported over a substantial period.<br />

Table 47: <strong>The</strong>rmoluminescence dating results. Precision shown is at confidence limits<br />

<strong>of</strong> one standard deviation.<br />

Lab# Site Sherd Grain size Date (ka) Date after<br />

fading<br />

correction<br />

(ka)<br />

uw471 Gharanga GH.183 fine 0.984±0.166 1.608±1.043<br />

uw472 Paniavile C.49 coarse 1.095±0.364 2.079±1.202<br />

uw473 Paniavile R379a4 coarse 1.801±0.296<br />

uw474 Paniavile P.165 coarse 0.790±0.363 0.950±0.495<br />

uw475 Honiavasa HV.5.120 coarse 1.903±0.290<br />

uw476 Honiavasa HV.1.48 coarse 1.903±0.290 2.339±0.750<br />

uw477 Miho MH.61 fine 1.798±0.278<br />

uw478 Zangana B.1 coarse 1.817±0.338<br />

464


UW474 suggests (at two standard deviations) that firing or use <strong>of</strong> quartz-calcite pottery<br />

continued until after 1940 BP (BP=before 2001 for this TL data) or in calendrical years,<br />

61AD. In combination with the radiocarbon evidence from NZA-12353 (Figure 140)<br />

UW474 suggest a minimum total temporal span for the intertidal-zone sites from 550BC<br />

to 61AD, or 611 years.<br />

Seriation:<br />

<strong>The</strong> technique chosen for seriation was correspondence analysis (CA), using the<br />

WINBASP program developed by Bonn <strong>University</strong>. Correspondence analysis will only<br />

produce a temporal seriation if the types/attributes in terms <strong>of</strong> which the units are being<br />

described are chronologically sensitive. “<strong>The</strong>re must be reasons to believe this in the first<br />

Table 48: Definitions <strong>of</strong> attribute codes used in seriation tables and plots.<br />

N_A_B_NB Neck, applied decoration, being a band, <strong>of</strong> nubbins<br />

L_F_BOPC Lip, fingernail impression, being a band, <strong>of</strong> opposed pinching<br />

N_F_B_PC Neck, fingernail impression, being a band, <strong>of</strong> opposed pinching<br />

F_MB__PC Fingernail impression, multiple bands, <strong>of</strong> opposed pinching<br />

L_I___BI Lip, impression, band <strong>of</strong>, inner edge <strong>of</strong> lip<br />

L_I___BO Lip, impression, being a band <strong>of</strong>, outer edge <strong>of</strong> lip<br />

L_I___BT Lip, impression, being a band <strong>of</strong>, top <strong>of</strong> lip<br />

L_I___BB Lip, impression, being a band <strong>of</strong>, both edges<br />

L_I__WAV Lip, impression, into a wave pattern (staggered inner and outer edges)<br />

L_D__WAV Lip, deformation, into a horizontal wave<br />

L_D__OCI Lip, deformation, discontinuous into individual spout-like impressions<br />

G_P____B anywhere on vessel (G=general), punctation, single horizontal band<br />

CLASS1LM Class 1 linear motif, anywhere on vessel<br />

R_CLS2LM Rim, class 2 linear motif<br />

S_CLS2LM Shoulder, class 2 linear motif<br />

STAMPING Anywhere on vessel, stamped decoration (including dentate)<br />

465


place (Shennan 1997:342).” <strong>The</strong> decorative attribute incidences (sherd counts) tabulated<br />

in Table 42 were used in the initial seriation reported below. To control for form variation,<br />

a number <strong>of</strong> these attributes were excluded from a second seriation, producing a different<br />

result to that <strong>of</strong> the total attribute set.<br />

In the seriations, attributes from Table 42 are coded as in Table 48. Schematic<br />

representation <strong>of</strong> attributes is given in Figure 143.<br />

Seriation 1: All Attributes, All Vessel Forms:<br />

Sample sizes by site for the various attributes are given in Table 42. Component 1 and<br />

Component 2 represent more than 84% <strong>of</strong> the total inertia, which reassures that the counts<br />

<strong>of</strong> the various attributes reduce well into two dimensions <strong>of</strong> variability (Table 49). <strong>The</strong> Qlt<br />

column <strong>of</strong> Table 50 indicates that some attributes are not well characterized by these two<br />

components (impressions on the outer lip and impressions on the top <strong>of</strong> the lip, have the<br />

lowest quality scores, while wave-impression <strong>of</strong> the lip, deformation <strong>of</strong> the lip by<br />

impression, and Class 2 linear motif on the shoulder score fairly low, suggesting these<br />

latter attributes are only moderately well characterized by the two summary components).<br />

<strong>The</strong> Mass column indicates that lip deformation into a wave, a band <strong>of</strong> pinching at the<br />

neck, and banded punctation have the greatest weight among the types by virtue <strong>of</strong> their<br />

greater incidence, while stamping, lip impression into a wave and a band <strong>of</strong> applied nubbins<br />

at the neck have the least sample-size-related weight <strong>of</strong> the types/attributes. <strong>The</strong>re are no<br />

outliers among the attribute contributions to inertia (Inr). Pinched band at the lip, Class 1<br />

linear motifs, band <strong>of</strong> punctation, and band <strong>of</strong> impressions on the inner lip have the greatest<br />

attribute contributions to inertia.<br />

<strong>The</strong> first Ctr column (Table 50) shows that applied band <strong>of</strong> nubbins at the neck<br />

(N_A_B_NB) contributes 8.0% <strong>of</strong> the inertia in Component 1 (80/1000), band <strong>of</strong> pinching<br />

at the lip (L_F_BOPC) contributes 22.5% <strong>of</strong> the Component 1 inertia (225/1000), etc.<br />

Thus we see that Component 1 is a product principally <strong>of</strong> three attribute incidences, these<br />

being: Class 1 linear motif incidence (CLASS1LM), band <strong>of</strong> pinching <strong>of</strong> the lip<br />

(L_F_BOPC), and a band <strong>of</strong> impressions on the inner lip (L_I___BI).<br />

466


Figure 143: Attributes used in seriations, groupings explained in <strong>chapter</strong> conclusions.<br />

467


Component 2 inertia is a result principally <strong>of</strong> the distribution through sites <strong>of</strong> bands <strong>of</strong><br />

punctation (G_P____B). Multiple bands <strong>of</strong> pinching, a band <strong>of</strong> pinching at the neck, and<br />

Class 2 linear motif on the shoulder are also important contributors to Component 2<br />

inertia. A plot <strong>of</strong> type (attribute) correspondence values for the two components is given<br />

in Figure 144.<br />

Table 49: Relative contribution <strong>of</strong> first and second components <strong>of</strong> CA using all<br />

attributes and forms.<br />

Component Iterations Norm Eigenvalue % Inertia Cumulative%<br />

1 28 0.051 0.583806 51.3 51.3<br />

2 8 0.082 0.376932 33.1 84.4<br />

Table 50: CA diagnostics by attribute, all attributes included.<br />

Attribute<br />

Qlt Mass Inr Comp1 Cor Ctr Comp2 Cor Ctr<br />

N_A_B_NB 971 20 42 -1534 962 80 -148 9 1<br />

L_F_BOPC 969 50 127 -1616 909 225 415 60 23<br />

N_F_B_PC 921 124 76 508 372 55 616 549 125<br />

F_MB__PC 854 50 63 298 62 8 -1062 791 151<br />

L_I___BI 982 61 94 -1307 977 179 -92 5 1<br />

L_I___BO 63 63 15 74 20 1 -106 43 2<br />

L_I___BT 0 61 20 0 0 0 -9 0 0<br />

L_I___BB 893 29 32 -195 30 2 -1043 863 83<br />

L_I__WAV 421 14 44 327 30 3 -1172 391 52<br />

L_D__WAV 669 156 42 409 542 45 198 127 16<br />

L_D__OCI 369 70 23 339 304 14 157 65 5<br />

G_P____B 951 122 136 299 70 19 -1056 881 362<br />

CLASS1LM 983 31 120 -2071 959 225 327 24 9<br />

R_CLS2LM 895 92 76 593 371 55 704 524 121<br />

S_CLS2LM 563 43 39 427 177 13 632 387 46<br />

STAMPING 840 13 49 -1899 815 78 328 24 4<br />

468


Figure 144: Correspondence plot (attributes) using all attributes and forms.<br />

<strong>The</strong> same sort <strong>of</strong> diagnostic statistics are given for the sites (Table 51). <strong>The</strong> two summary<br />

components give a good representation <strong>of</strong> the variability at Honiavasa (999/1000).<br />

Similarly, but to a lesser extent, the variability at all other sites except the three with the<br />

least mass (small sample sizes) is well represented. <strong>The</strong> Honiavasa site is an outlier in terms<br />

<strong>of</strong> its contribution to total inertia, contributing more than 80% <strong>of</strong> Component 1 inertia.<br />

Gharanga makes the greatest contribution to Component 2 inertia, with Hoghoi, Miho and<br />

Zangana South making significant contributions to Component 2 also. Honiavasa’s<br />

contribution to Component 2 inertia is negligible (1.8%).<br />

469


Table 51: CA diagnostics by site, all attributes included (*=inertia outlier).<br />

SITES Qlt Mass Inr Comp1 Cor Ctr Comp2 Cor Ctr<br />

PANIAVIL 703 187 70 348 284 39 422 418 88<br />

HOGHOI__ 817 165 87 174 50 9 -676 767 201<br />

MIHO____ 794 180 98 507 416 79 484 378 112<br />

HONIAVAS 999 167 418* -1674 985 803 201 14 18<br />

GHARANGA 866 101 164 158 13 4 -1255 852 421<br />

NUSAROVI 39 20 24 -112 9 0 203 30 2<br />

KOPO____ 154 4 15 417 38 1 -733 116 5<br />

ZANNORTH 487 61 41 327 140 11 -516 348 43<br />

ZANSOUTH 761 115 84 520 327 53 599 434 109<br />

A plot <strong>of</strong> these two components is given in Figure 145. <strong>The</strong> correspondence scores in the<br />

Cor columns provide the coordinates for these plots. Honiavasa is an outlier along the<br />

Component 1 axis, the most similar site along Component 1 being Nusa Roviana, while the<br />

other sites separate into two groups at opposing ends <strong>of</strong> Component 2. Nusa Roviana has<br />

one example <strong>of</strong> a Class 1 linear motif (sherd NR34), but is closer to Honiavasa more by<br />

default, by absence <strong>of</strong> characteristics, leading to a location on the plot close to the average,<br />

or 0,0 position. <strong>The</strong> single occurrence <strong>of</strong> stamping at Nusa Roviana will also have had a<br />

slight effect on the plot. It must be remembered that Nusa Roviana was not well<br />

characterized by these two components. Stamping and Class 1 linear motif occurred on the<br />

same sherd, so the similarity between Nusa Roviana and Honiavasa hinges on a single<br />

sherd, and thus should not be made too much <strong>of</strong>.<br />

470


Figure 145: Correspondence plot (sites) using all attributes and forms.<br />

As discussed in the form and decoration <strong>chapter</strong>s (Chapters 8 and 9) there are some<br />

decorative attributes that are highly correlated with aspects <strong>of</strong> vessel form variation, and<br />

which may be decoration specific to vessels <strong>of</strong> a particular function. In order to minimise<br />

the risk <strong>of</strong> incorrectly seriating contemporaneous functional variation, these attributes were<br />

excluded from analysis below. Excluded were: Class 1 linear motifs; bands <strong>of</strong> applied<br />

nubbins at the neck; and stamping; all highly correlated with carinated vessel forms.<br />

Multiple bands <strong>of</strong> fingernail pinching were also excluded due to occurrence only on short-<br />

rim heavily excurvate Gharanga-type vessels.<br />

Also, some lip impression similarities between Gharanga form/decoration type and<br />

some Honiavasa tall excurvate rims are likely to be analogous rather than homologous<br />

similarity (see Chapter 9), and banded lip impressions are thus excluded from subsequent<br />

471


seriations, with the exception <strong>of</strong> lip impressions on both edges <strong>of</strong> the lip forming a wave<br />

pattern; this decorative attribute occurred mainly in the Gharanga site on tall weakly<br />

excurvate rims within the “general purpose” Form 6c/2c category, and indications from<br />

this pattern <strong>of</strong> occurrence are a relatively restricted temporal occurrence.<br />

Seriation 2: Form-correlated Attributes Removed:<br />

Cumulative inertia <strong>of</strong> the first two components is 88%, suggesting that the patterning in<br />

the data summarizes well into these components (Table 52). <strong>The</strong> three components<br />

together produce an average quality <strong>of</strong> characterization for the attributes <strong>of</strong> 81%<br />

(812/1000) (Table 52), with three attributes only moderately characterized (wave-<br />

deformation <strong>of</strong> the lip, discontinuous finger-deformation <strong>of</strong> the lip and Class 2 linear motif<br />

on the shoulder). Of these, Class 2 linear motif on the shoulder has a low mass figure,<br />

related to small sample size. Component 1 contributions to inertia are almost wholly<br />

dominated by fingernail pinch band at the lip, while the major contributor to Component<br />

2 inertia is banded punctation (G_P___B).<br />

Class 2 linear motif on the rim and a band <strong>of</strong> pinching at the neck make significant<br />

contributions to Component 2 inertia also, while the reasonably numerous wave<br />

deformation <strong>of</strong> the lip and discontinuous finger-deformation <strong>of</strong> the lip classes make very<br />

little contribution to the inertia <strong>of</strong> either <strong>of</strong> the first two components. This is particularly<br />

interesting, as these ubiquitous decorative techniques may indicate either a long production<br />

span, or possibly a temporally central production span within the Roviana intertidal milieu,<br />

where occupation spans <strong>of</strong> sites tend to incorporate discard <strong>of</strong> these attributes by virtue<br />

<strong>of</strong> their temporal centrality. Component three may hold the key in this regard, because<br />

these attributes, and particularly the related technique <strong>of</strong> lip impression into a wave, have<br />

a lot <strong>of</strong> inertia there. <strong>The</strong> correspondence scores <strong>of</strong> these attributes for Components 2 and<br />

3 are plotted in Figure 146.<br />

<strong>The</strong> position <strong>of</strong> pinching <strong>of</strong> the lip on this plot can be disregarded as it is an outlier<br />

472


in Component 1, which is not shown. Lip impression into a wave pattern is also an outlier,<br />

in the low-inertia Component 3, but small sample size must be born in mind here. Although<br />

most examples occur in a single site, absence from others might be a sample size effect.<br />

<strong>The</strong> tight grouping <strong>of</strong> neck pinching, Class 2 linear motif on the rim, and Class 2 linear<br />

motif on the shoulder confirms the “Miho” decorative style as a relatively coherent entity,<br />

as suggested in Chapter 9. <strong>The</strong> separation <strong>of</strong> lip deformation into a wave and discontinuous<br />

finger deformation <strong>of</strong> the lip from the “Miho style” grouping results largely from the spatial<br />

separation <strong>of</strong> these attributes within the greater Zangana site, although wave-deformation<br />

<strong>of</strong> the lip occurs in almost all sites to some extent, which is why these attributes lie close<br />

to the intersection <strong>of</strong> the axes. Overall it is reassuring that the attributes order in the same<br />

way in Component 2 as they did in Figure 144 despite omission <strong>of</strong> some potential problem<br />

attributes.<br />

Table 52: Eigenvalues and contributions to intertia <strong>of</strong> CA components, for data<br />

excluding form-correlated attributes.<br />

Component Iterations Norm Eigenvalue % Inertia Cumulative<br />

%<br />

1 28 0.051 0.611850 53 53.0<br />

2 9 0.004 0.407938 35.3 88.3<br />

3 10 0.041 0.080177 6.9 95.2<br />

Table 53: CA diagnostic table for attributes, form-correlated attributes omitted (*=intertia<br />

outlier).<br />

Name Qlt Mass Inr Comp1 Cor Ctr Comp2 Cor Ctr Comp3 Cor Ctr<br />

L_F_BOPC 1000 75 479* 2676 969 876 479 31 42 -21 0 0<br />

N_F_B_PC 959 184 56 -111 35 4 -547 850 135 -161 74 60<br />

L_I__WAV 972 21 81 -591 80 12 1255 359 83 -1530 533 624<br />

L_D__WAV 638 233 26 -165 213 10 -133 138 10 192 288 107<br />

L_D__OCI 485 104 18 -90 39 1 -62 19 1 295 426 113<br />

G_P____B 996 182 240 -551 199 90 1099 792 538 89 5 18<br />

R_CLS2LM 862 136 64 -158 46 6 -663 809 147 -62 7 7<br />

S_CLS2LM 586 64 35 36 2 0 -525 441 43 -298 142 71<br />

473


Turning to the table <strong>of</strong> scores for sites (Table 54) it is clear that all sites except Kopo and<br />

Nusa Roviana are well characterized by the three components <strong>of</strong> the analysis. <strong>The</strong> mass<br />

column indicates that these latter sites had small sample sizes, and were given little weight<br />

in the calculation <strong>of</strong> inertia values. Honiavasa site is still an outlier in its contribution to<br />

total inertia, by virtue <strong>of</strong> a contribution <strong>of</strong> almost 87% <strong>of</strong> Component 1 inertia, which, as<br />

can be seen in Figure 146, is mostly to do with the incidence <strong>of</strong> a band <strong>of</strong> pinching at the<br />

lip. Gharanga site dominates the contributions to Component 2 inertia, but Hoghoi, Miho<br />

and Zangana South are important also, the first two by virtue <strong>of</strong> the high incidence <strong>of</strong><br />

punctation, the last because <strong>of</strong> neck pinching and Class 2 linear motif on the rim. A plot<br />

<strong>of</strong> the first two component correspondence scores is given in Figure 147.<br />

Figure 146: Correspondence plot (attributes) excluding form-correlated attributes<br />

474


Table 54: CA diagnostics by site, form-correlated attributes omitted (*=inertia outlier).<br />

SITE Qlt Mas<br />

s<br />

Inr Comp1 Cor Ctr Comp<br />

2<br />

475<br />

Cor Ctr Comp<br />

3<br />

Cor Ctr<br />

PANIAVIL 840 230 40 34 6 0 -410 834 95 14 1 1<br />

HOGHOI__ 948 144 104 -439 231 45 727 634 187 264 84 126<br />

MIHO____ 822 225 71 -192 101 14 -502 691 139 104 30 30<br />

HONIAVAS 1000 75 476* 2661 965 86<br />

7<br />

508 35 47 6 0 0<br />

GHARANGA 989 80 180 -609 143 49 1358 710 362 -595 136 354<br />

NUSAROVI 325 19 9 -248 114 2 -102 19 0 322 192 24<br />

KOPO____ 464 5 13 -410 61 1 811 237 9 678 166 31<br />

ZANNORTH 872 70 44 -409 230 19 528 382 47 435 260 164<br />

ZANSOUTH 957 152 63 -107 24 3 -551 635 113 -377 297 270<br />

In comparison to Figure 145, Figure 147 shows a different ordering <strong>of</strong> sites in<br />

Component 1.<br />

Figure 147: Correspondence plot (sites) excluding form-correlated<br />

attributes.


While this might seem relatively minor in view <strong>of</strong> the stability <strong>of</strong> Component 2 ordering<br />

between these two plots, Component 1 is crucial to the overall relative dating <strong>of</strong> the three<br />

broad styles that make up the series (the Honiavasa sample, Miho style and Gharanga<br />

style) as it is Component 1 which shows which <strong>of</strong> the site assemblages is more similar to<br />

the Honiavasa Lapita site. Where previously Nusa Roviana, Gharanga and Hoghoi were<br />

closest to Honiavasa in terms <strong>of</strong> Component 1, now Paniavile is closest, by virtue <strong>of</strong> the<br />

presence <strong>of</strong> five examples <strong>of</strong> opposed-pinch band on the outer lip there. Zangana South<br />

is next closest due to a single occurrence <strong>of</strong> the same attribute there. <strong>The</strong>se sample sizes<br />

show how tenuous the homologous links between Honiavasa and the other sites become,<br />

when the (potentially analogous) lip impression similarity is excluded. <strong>The</strong>re are echoes in<br />

this situation <strong>of</strong> the dilemma faced by Specht in deciding whether the transition from Buka<br />

phase to Sohano phase on Bougainville was gradual, but poorly sampled, or whether<br />

cultural replacement had occurred. It seems clear in the present case that the Honiavasa<br />

site is a site largely alone, with the exception <strong>of</strong> occasional sherds with similar attributes<br />

occurring elsewhere, and that the current sample is insufficient to address this question<br />

from ceramic stylistic evidence alone.<br />

With the omission <strong>of</strong> lip impressions Gharanga is now the most distant site in<br />

Component 1, and all the sites in which a single band <strong>of</strong> punctation is dominant are now<br />

at the opposite end <strong>of</strong> Component 1 to the Honiavasa site. Ordering in Component 1 <strong>of</strong><br />

these other sites roughly mirrors ordering in Component 2, with the group nearest to<br />

Honiavasa in Component 1 (Paniavile, Miho, Zangana South and Nusa Roviana) forming<br />

a separate group in Component 2 also.<br />

Seriation 3: Honiavasa Excluded and Form-correlated Attributes Excluded:<br />

If Honiavasa, the outlier, is excluded from the data, some <strong>of</strong> the minor components might<br />

be expected to assume greater weight in the CA, which is not the case (Table 57, Table<br />

55, Table 56 and Figure 148).<br />

476


Table 57: CA eigenvalues and component contributions to inertia omitting Honiavasa<br />

data and form-correlated attributes<br />

Component Iterations Norm Eigenvalue % Inertia Cumulative<br />

1 8 0.024 0.426248 72 72.0<br />

2 21 0.048 0.081960 13.9 85.9<br />

3 15 0.048 0.046658 7.9 93.8<br />

Table 55: CA diagnostics for attributes, omitting Honiavasa data and form-correlated<br />

attributes.<br />

ATTRIBUTE Qlt Mas<br />

s<br />

Table 56: CA diagnostics by site, omitting Honiavasa data and form-correlated<br />

attributes.<br />

SITE Qlt Mass Inr Comp1 Cor Ctr Comp2 Cor Ctr Comp3 Cor Ctr<br />

PANIAVIL 978 249 122 441 673 114 27 2 2 296 302 465<br />

HOGHOI__ 947 156 188 -779 852 222 -249 87 118 75 8 19<br />

MIHO____ 986 243 117 411 594 96 -142 71 59 -302 321 475<br />

GHARANG<br />

A<br />

990 87 342 -1402 844 400 575 142 350 -105 5 20<br />

NUSAROVI 233 20 16 32 2 0 -325 230 26 16 1 0<br />

KOPO____ 472 6 25 -855 280 10 -680 177 33 -193 14 5<br />

ZANNORT<br />

H<br />

Inr Comp1 Cor Ctr Comp2 Cor Ctr Comp3 Cor Ctr<br />

L_F_BOPC 908 17 58 691 243 19 284 41 17 1106 623 455<br />

N_F_B_PC 944 197 95 494 855 112 145 74 50 -67 16 19<br />

L_I__WAV 981 23 156 -1300 424 92 1457 533 599 -311 24 48<br />

L_D__WAV 561 246 36 92 97 5 -184 390 101 80 74 34<br />

L_D__OCI 567 107 36 34 6 0 -313 487 128 -123 75 34<br />

G_P____B 996 197 444* -1151 991 610 -76 4 14 39 1 7<br />

R_CLS2LM 936 147 107 578 780 116 30 2 2 -256 153 208<br />

S_CLS2LM 877 66 69 539 473 45 332 179 89 371 224 196<br />

855 75 78 -590 566 61 -416 281 158 73 9 9<br />

ZANSOUTH 932 165 113 500 616 97 355 311 254 -45 5 7<br />

477


Only a single dimension <strong>of</strong> high inertia remains in the data, which is dominated by the<br />

inertia <strong>of</strong> bands <strong>of</strong> punctation, and it produces groupings much like Component 2 in the<br />

previous plots, with Gharanga at one extreme (tailed by Kopo, with a tiny sample, and<br />

Hoghoi, then Zangana North). Zangana South is at the opposite extreme, with Paniavile<br />

and Miho nearby, and Nusa Roviana tending in the same direction, although sample size<br />

for that site is negligible. <strong>The</strong> second component has only 14% inertia, and mainly involves<br />

the count <strong>of</strong> impression <strong>of</strong> the lip into a wave pattern, which occurs five times at Gharanga<br />

and twice at Zangana South.<br />

Discussion <strong>of</strong> Correspondence Analyses: the Alternatives:<br />

<strong>The</strong> search for temporal variability centered on the exclusion <strong>of</strong> form-correlated decorative<br />

attributes. Exclusion <strong>of</strong> CLASS1LM and STAMPING made no difference to the<br />

Figure 148: Correspondence plot (sites), Honiavasa sample and form-correllated<br />

attributes omitted from data-set.<br />

478


ordination <strong>of</strong> sites or attributes, with the exception <strong>of</strong> Nusa Roviana, a small site sample<br />

with a single sherd on which two <strong>of</strong> the excluded attributes occurred (Class 1 linear motifs<br />

and stamping). <strong>The</strong> excluded sherd tended to make Nusa Roviana more similar to<br />

Honiavasa, the latter site being an outlier in all analyses.<br />

<br />

<strong>The</strong> effect on the seriation <strong>of</strong> excluding bands <strong>of</strong> lip impression is worthy <strong>of</strong> note.<br />

If these analysis are interpreted in temporal terms, then two temporal ordinations <strong>of</strong><br />

groupings in the data (groupings are summarized in Table 58) can be constructed, either<br />

A-B-C or A-C-B (or a mirror image <strong>of</strong> either <strong>of</strong> these, as seriation does not specify a<br />

direction for time). <strong>The</strong>se two alternatives, when combined with the notion that the<br />

Honiavasa (A) phase is most Lapita-like, and therefore earliest (for which there is some<br />

supporting evidence from TL), yield two alternative developmental modes. If Honiavasa<br />

Table 58: Summary <strong>of</strong> variability<br />

Phase<br />

Symbol<br />

A<br />

B<br />

C<br />

Multipurpose vessel<br />

decoration<br />

<br />

one-piece construction PINCHED<br />

BAND LIPS, IMPRESSED LIPS<br />

(especially the inner lip), some<br />

coarse wave deformation <strong>of</strong> the lip<br />

One-piece construction PINCHED<br />

BAND NECKS, CLASS 2<br />

LINEAR MOTIF RIMS, CLASS 1<br />

LINEAR MOTIF SHOULDERS,<br />

tall excurvate rim form with thin<br />

fragile lips, thick robust necks.<br />

Tendency towards wavedeformation<br />

<strong>of</strong> the lip, when<br />

preserved.<br />

One-piece construction<br />

PUNCTATE BAND in the vicinity<br />

if the neck, IMPRESSED LIPS,<br />

some lip impression into a wave<br />

pattern<br />

479<br />

Temporally associated<br />

decorative attributes and<br />

vessel forms/functions<br />

Robust Slab-Constructed carinated<br />

vessels with class 1 linear motifs,<br />

stamped decoration, and band <strong>of</strong><br />

applied nubbins at the neck, some<br />

compound or collar rims, some<br />

evidence for shallow bowls and robust<br />

storage vessels with low breakage<br />

rates.<br />

Various small bowls or flared rims,<br />

low discard rate <strong>of</strong> large robust vessels<br />

(storage vessels?)<br />

Gharanga short-rim variant, thinwalled<br />

in many cases, with multiple<br />

bands <strong>of</strong> fingernail pinching on the<br />

shoulder and bands <strong>of</strong> lip impression<br />

(especially the inner lip), occasional<br />

discard <strong>of</strong> punctate bowls.


is early, Miho style (B) is intermediate and Gharanga style(C) is late then the period over<br />

which discard occurred is characterized by a change from slab-constructed forms to one-<br />

piece forms, initially using a tall rim similar to that used in slab construction, but eventually<br />

replacing this for some purposes with a more rugged short, strongly outcurved rim form<br />

that survives well in the swash zone in spite <strong>of</strong> very thin construction in some cases.<br />

Alternatively, if Gharanga/Kopo style is earlier than Miho style, then the transition to thin,<br />

short-rim pots from Honiavasa heavy excurvate-rimmed pots is more rapid, the loss <strong>of</strong> the<br />

heavily carinated Class 2 decorated pots or jars is also relatively rapid, and by the terminal<br />

stage, Miho-style linear motifs (unbounded this time) reappear, along with tall rims arising<br />

from Kopo-style tall rimmed punctate-decorated vessels.<br />

Conclusions: Integrating 14 C, TL and Seriation:<br />

14 C data suggested we can be reasonably confident that the Roviana intertidal sites have<br />

a total occupation span <strong>of</strong> at least 160 years, unless there is an old-wood problem with the<br />

smoke-derived date from Hoghoi. It was not possible to associate either <strong>of</strong> these dates<br />

with a particular style <strong>of</strong> ceramic production, although the thin-necked pot from Hoghoi<br />

with a tall everted rim seems intermediate in some attributes between the Gharanga style<br />

and either Honiavasa plain pottery or Miho-style pottery.<br />

<strong>The</strong> addition <strong>of</strong> eight TL dates expands the minimum occupation span <strong>of</strong> the<br />

intertidal sites to 611 years. Also, if the TL ages from quartz-calcite hybrid tempered<br />

sherds uncorrected for fading are taken as an ordinal-scale indication <strong>of</strong> age ( Dickinson’s<br />

conclusion was that these tempers all originate from a single coastline, which might<br />

provide some confidence that, although they fade anomalously, they may all fade in the<br />

same anomalous way), then the Honiavasa site is likely to be oldest, while some <strong>of</strong> the<br />

pottery from Miho, Paniavile and Zangana South is mostly intermediate in age between<br />

Honiavasa and Gharanga. Other pottery from Gharanga and Paniavile is younger still.<br />

More TL dates are necessary to test this argument, and could pr<strong>of</strong>itably use some <strong>of</strong> the<br />

decorated Q-C tempered sherds held in the collections at present, especially sherd MH033,<br />

480


decorated with a band <strong>of</strong> punctation at the neck.<br />

styles, either:<br />

or:<br />

Seriation analyses can be summarized as producing two different sequences <strong>of</strong><br />

HoniavasaMihoGharanga/Kopo<br />

HoniavasaGharanga/KopoMiho<br />

(the seriation was more fine-grained than this, being attribute-based, but the major<br />

chronological issue can be phrased in terms <strong>of</strong> styles). TL results suggest that the<br />

Honiavasa site is older than the others, and that the Miho style predates Gharanga/Kopo,<br />

although more evidence is needed to test this. This conclusion rests on an assumption that<br />

the feldspar grains in quartz-calcite hybrid tempers used in TL dating all fade in a similar<br />

manner, allowing a relative chronology.<br />

<strong>The</strong> seriation analysis tended to show groups <strong>of</strong> attributes clustering into styles,<br />

which suggested the homologous links between Honiavasa site and the other sites were<br />

few (but present), which is most economically explained as resulting from historical<br />

discontinuity in the sample. Also, the phyletic links that could show the relationship<br />

between Miho and Gharanga/Kopo styles in detail were largely absent (but not completely<br />

absent). This again is interpreted as a historical gap in the sample, rather than indicating<br />

any essential, immutable reality to the Miho and Gharanga types. In the absence <strong>of</strong><br />

evidence to the contrary, any conclusion that these data clusters represent cultural<br />

replacement <strong>of</strong> any sort is unwarranted. As discussed in the next <strong>chapter</strong>, there is non-<br />

ceramic evidence for heritable continuity across the sample. If the corpus <strong>of</strong> sites were<br />

much larger but these discontinuities persisted, then explanation for the discrete styles<br />

might be sought in historical events. <strong>The</strong> current state <strong>of</strong> sampling <strong>of</strong> the New Georgia<br />

group does not rule out gradual continuous changes in some form and decoration<br />

attributes over time for the larger region. Whether such changes all happened at Roviana<br />

Lagoon, but outside the current site sample, or whether the action moved <strong>of</strong>fstage, beyond<br />

the sampling frame, at some times, is unknown from the present evidence.<br />

481


482


Introduction:<br />

CHAPTER 13:<br />

SUMMARY AND CONCLUSIONS<br />

<strong>The</strong> key question laid out in Chapter One was whether there was a gap in the past<br />

distribution <strong>of</strong> Lapita, consistent with an avoidance or leap-frogging colonization model,<br />

or whether the near-oceanic Solomon Islands, as represented by New Georgia, were<br />

settled early in the Lapita ceramic series. As the early Roviana ceramics are all in the sea<br />

there were special questions concerning formation processes <strong>of</strong> this particular record, and<br />

also a need for adaptation <strong>of</strong> archaeological method in some respects to fit this<br />

archaeological landscape.<br />

<strong>The</strong> broad question concerning Lapita distribution across the region <strong>of</strong> Near<br />

Oceania was broken down into a number <strong>of</strong> more specific topics: how can the Roviana<br />

intertidal record be interpreted in behavioural terms (and to what extent is patterning due<br />

to natural rather than cultural formation processes)? How can a high-resolution chronology<br />

be constructed to allow fine chronological control and enable the historicist approaches<br />

that are so <strong>of</strong>ten attempted using inappropriately coarse radiocarbon chronologies? How<br />

much <strong>of</strong> a sample is needed, and what sort <strong>of</strong> sample? how can sample size be quantified?<br />

What potential is there for bias given particular sets <strong>of</strong> formation processes, and what are<br />

the likely biases present in the samples?<br />

In tackling these questions, conclusions <strong>of</strong> significance to resource management<br />

and future research were drawn regarding the fragility <strong>of</strong> these swash-zone sites, the<br />

poverty <strong>of</strong> information remaining, and extreme sensitivity in terms <strong>of</strong> information content<br />

to further removal <strong>of</strong> material. In spite <strong>of</strong> the apparent poverty <strong>of</strong> the record, the high<br />

archaeological visibility <strong>of</strong> lag deposits <strong>of</strong> pottery and lithics means there is unusually<br />

483


good potential for fine-grained chrono-stylistic studies.<br />

Summary:<br />

A review <strong>of</strong> approaches to ceramic classification and seriation method in Chapter<br />

1 identified a need for materialist approaches to artifact classification capable <strong>of</strong> capturing<br />

stylistic drift, compatible with a descent-with-modification mode <strong>of</strong> ceramic change, in<br />

order to develop fine-grained stylistic chronology. <strong>The</strong> review identified a need to control<br />

for functional variability, as a means <strong>of</strong> isolating temporal variability.<br />

A review <strong>of</strong> theories <strong>of</strong> form-function correlation found that functional classes were<br />

best kept broad, and that use-life theory predicts that general purpose vessels that included<br />

a cooking function should dominate vessel samples. Seriation should focus on forms that<br />

fell into this broad use-category, and on decoration that cross-cut form-function classes.<br />

<strong>The</strong> review identified a disjuncture between descriptive/classificatory units for<br />

Lapita and post-Lapita (Mead’s system or Anson’s system principally for Lapita, and the<br />

Frost-Irwin attribute combination approach for later pottery). In order to span the Lapita-<br />

post-Lapita period these differences needed to be resolved. <strong>The</strong> relative simplicity <strong>of</strong><br />

decoration on utilitarian pottery, which dominates post-Lapita assemblages at least, meant<br />

that descriptive and classificatory schemes needed to be sensitive to slight differences in<br />

decoration. Also, in relation to units <strong>of</strong> quantification, there is a need for measures <strong>of</strong><br />

attribute frequency and sample sizes that are not biased by differences in assemblage<br />

brokenness, or differential preservation <strong>of</strong> parts <strong>of</strong> the vessel. Accordingly, it was<br />

determined that units <strong>of</strong> description and classification should pay close attention to<br />

location on the vessel, and should focus on those parts <strong>of</strong> the vessel that preserve well and<br />

are easily identified even as small fragments, allowing the structure <strong>of</strong> decoration across<br />

the vessel as represented by sherd samples to be captured.<br />

A review <strong>of</strong> temporal constructs for Lapita (Chapter 2) found that temporal<br />

484


esolution <strong>of</strong> the available chronologies was alarmingly similar to the various definitions<br />

<strong>of</strong> Lapita, and could be broadly characterized as culture-historical in resolution, with<br />

Lapita continuing to have the status <strong>of</strong> a secure temporal horizon, but with potential in<br />

current higher-resolution constructions <strong>of</strong> Lapita temporal variability for mis-assignment<br />

<strong>of</strong> other dimensions <strong>of</strong> variability to time. This is partly a result a lack <strong>of</strong> clear sample<br />

evaluation, an important step, particularly when constructing occurrence seriations and<br />

interpreting motif-sharing. <strong>The</strong>se factors undermine the security <strong>of</strong> current constructions<br />

<strong>of</strong> the “Lapita Ceramic Series”.<br />

While temporal constructs are seen as secure by several investigators, for example,<br />

Anson's “Early Far Western/Western” distinction (Anson 1983, 1986, 1987, 1990, 2000),<br />

Summerhayes’ Early/Middle/Late universal Lapita series (Summerhayes 2000a, 2001,<br />

2002), and Sand’s Southern Lapita series for New Caledonia (Sand 1997a, 1999, 2000,<br />

2001), there is a tendency for site samples to be slotted in to these schemes without much<br />

control for functional variation especially, and geographic variation to a lesser extent.<br />

While I would not suggest these constructions are necessarily wrong, these are seen here<br />

as important works in progress rather than a secure foundation for comparative dating <strong>of</strong><br />

the Roviana materials.<br />

While Best’s sequence at Lakeba was judged to be a more secure stratigraphic<br />

demonstration <strong>of</strong> temporal changes from Lapita to other styles, uncertainties remain, and<br />

the Eastern chronology need not apply elsewhere. Constructing a phyletic seriation to<br />

argue that it does, as Spriggs does in “<strong>The</strong> Changing Face <strong>of</strong> Lapita” (Spriggs 1990) and<br />

as Best does in “A View from the East” (Best 2002), is putting the cart before the horse.<br />

Secure regional sequences need to be compared and cross-matched to assess whether there<br />

are universal Lapita-wide parallel changes before doing this. It cannot be assumed that<br />

there are such changes, to justify picking a phyletic series from across vast swaths <strong>of</strong> the<br />

Pacific, because this assumes the answer to a primary question to be asked <strong>of</strong> the material.<br />

In this final <strong>chapter</strong> some <strong>of</strong> the difficulties encountered in trying to reconcile the Roviana<br />

485


early ceramics with other temporal constructs will be elaborated.<br />

An increased emphasis on oceanography (wave processes) and the consequences<br />

<strong>of</strong> these for archaeological preservation and probability <strong>of</strong> detection was advocated in<br />

Chapter 3. Previous survey method and results in near Oceania were reviewed, and the<br />

results <strong>of</strong> two surveys <strong>of</strong> the Roviana Lagoon region were presented in terms <strong>of</strong> a sample-<br />

surveying approach. For Lapita, the informant prospection survey (Roviana survey)<br />

obtained similar results to other informant-prospection surveys in the Near-Oceanic<br />

Solomon Islands, i.e. no hits (other than a single sherd), in spite <strong>of</strong> informants volunteering<br />

the locations <strong>of</strong> a number <strong>of</strong> post-Lapita or Lapita-derived intertidal sites. More intensive<br />

intertidal coverage <strong>of</strong> the complex coastline <strong>of</strong> the more restricted Kaliquongu area located<br />

a single Lapita site.<br />

Although a sample <strong>of</strong> one suggests attribution <strong>of</strong> different results to different<br />

survey methods should be made with caution, I think we have a better idea, as a result <strong>of</strong><br />

the Kaliquongu survey, how to locate sites <strong>of</strong> the Lapita period in the New Georgia area.<br />

Despite the small site-sample size for Lapita, Lapita recorded site density (as opposed to<br />

past site density) was well within the range recorded elsewhere in Near Oceania, <strong>of</strong>fering<br />

some evidence in support <strong>of</strong> a model <strong>of</strong> rapid and comprehensive spread <strong>of</strong> Lapita rather<br />

than leap-frog colonization or avoidance modes. Given that the Kaliquongu survey covered<br />

only a small fraction <strong>of</strong> the New Georgia intertidal zone and only one major reef passage<br />

into the chain <strong>of</strong> landlocked lagoons encircling New Georgia, it is highly likely that<br />

temporal-stylistic diversity will increase with further survey <strong>of</strong> similar locations around<br />

New Georgia, which, if this turns out on further analysis to be the case, would in turn<br />

strengthen support for a model <strong>of</strong> early, rapid spread. While speculative, this <strong>of</strong>fers a<br />

testable hypothesis for future work.<br />

What is clear from the Kaliquongu results is that we would need a lot more<br />

evidence than currently in hand to rule out early Lapita settlement <strong>of</strong> the New Georgia<br />

group. Furthermore, given the results <strong>of</strong> the review <strong>of</strong> the Lapita ceramic series, it is<br />

486


difficult to say on current evidence just how old the Honiavasa materials are, although they<br />

suggest a sufficient period <strong>of</strong> divergence from an early-Lapita ancestor to lack pedestalled<br />

bowls and flat-based dishes (although absence from the current sample could be a sampling<br />

error given the small sample <strong>of</strong> slab-built and carinated Lapita forms currently recorded<br />

from Roviana Lagoon).<br />

Chapter 4 gave details <strong>of</strong> the ceramic database, and coding <strong>of</strong> units <strong>of</strong> ceramic<br />

description in the core files on the appended data CD. Core sherd attributes <strong>of</strong> size, weight<br />

and fabric were recorded in the “Master.db” table, while thickness, form and decoration<br />

were coded in separated tables organizing multiple records by vessel part for each sherd.<br />

This is a departure from the “diagnostic sherds’ approach where analysis focussed only on<br />

a small subset <strong>of</strong> sherds which met (usually unstated) criteria. <strong>The</strong> approach allows close<br />

control for part representation when querying the data, and allows convenient access to<br />

a large set <strong>of</strong> attributes through a form layout <strong>of</strong> master record and client tables. Salient<br />

attributes from this relational database were reconstructed into a single flat table for spatial<br />

analysis and seriation.<br />

Fabric classes included several placering variants <strong>of</strong> volcanic sands, most <strong>of</strong> which<br />

are thought to be local, and an exotic quartz-calcite hybrid temper, thought on current<br />

evidence to be from a continental location to the west, but other less exotic sources cannot<br />

be ruled out on current evidence. <strong>The</strong>re were few calcite-only tempered sherds. <strong>The</strong><br />

quartz-calcite tempers were much more quartzose than any quartz-calcite hybrid tempers<br />

from the Bismarck Archipelago, with ratio <strong>of</strong> coarse K-feldspar-to-Plagioclase suggestive<br />

<strong>of</strong> granitic origin. This raises the prospect <strong>of</strong> low frequency <strong>of</strong> trans-Solomon Sea or trans-<br />

Coral-Sea pottery transfer, which would accord well with some <strong>of</strong> the materials from<br />

Reef-Santa-Cruz Lapita sites, and with Irwin’s voyaging models, but is anomalous in that<br />

no other sites in Near Oceania are known to have yielded similar pottery temper sand<br />

evidence (Felgate & Dickinson 2001). <strong>The</strong> explanation preferred by Felgate and<br />

Dickinson is thus unlikely to find much favour with the wider Pacific archaeological<br />

487


audience unless such a source <strong>of</strong> pottery is physically located, which is beyond the scope<br />

<strong>of</strong> the current research, so the implications <strong>of</strong> this temper remain an open question.<br />

Form <strong>of</strong> vessel parts was recorded with the aim <strong>of</strong> providing sufficient information<br />

to enable a reasonably accurate depiction <strong>of</strong> vessel shape variability. Nominal<br />

morphological classes such as everted rim, restricted neck, carination, etc were thus<br />

supplemented with a series <strong>of</strong> orientation, curvature and distance measurements. <strong>The</strong>se<br />

together with thickness measurements at a series <strong>of</strong> defined points enable more traditional<br />

nominal classes <strong>of</strong> vessel form to be deconstructed for the purpose <strong>of</strong> an exploratory<br />

analysis <strong>of</strong> variability, in tune with the materialist, evolutionist approach to identification<br />

<strong>of</strong> stylistic and functional variation.<br />

Similarly, decoration was coded by vessel part, employing formulaic descriptions<br />

<strong>of</strong> the patterned layout <strong>of</strong> elements for simple banded decorations, and reference to<br />

drawings for more complex or larger, or fragmentary designs. <strong>The</strong>re was a conscious effort<br />

in coming up with this scheme to separate decorative technique from decorative element,<br />

and from the layout <strong>of</strong> elements. Metric variability <strong>of</strong> formulaic banded designs was<br />

recorded, some <strong>of</strong> which information turned out to be salient to some <strong>of</strong> the basic research<br />

questions when analysed in Chapter 8.<br />

Chapter 5 provided a detailed investigation <strong>of</strong> vessel brokenness and completeness,<br />

with several practical outcomes important to the conclusions <strong>of</strong> the thesis. While many <strong>of</strong><br />

the samples comprised quite large sherds, vessel completeness as evidenced by<br />

construction <strong>of</strong> lip-sherd vessel families was extremely low, with mostly singleton lip<br />

sherds present. A total breakage population was estimated by constructing a virtual<br />

assemblage <strong>of</strong> whole rims broken in the manner <strong>of</strong> the sample, and iteratively sampling it<br />

at various levels to see what range <strong>of</strong> sampling fractions yielded completeness data similar<br />

to that <strong>of</strong> the real sample (Felgate & Bickler n.d). A statistical approach using the same<br />

completeness data and a “number <strong>of</strong> species” algorithm obtained similar results to the<br />

simulation, but the simulation results were preferred as more precise due to the use<br />

488


<strong>of</strong> brokenness information.<br />

Around 99% <strong>of</strong> the sherdage <strong>of</strong> the original breakage population has not made it<br />

into the sample, and the breakage population vessel count would have been at least five<br />

times larger than the vessel count represented in the lip sample. As little <strong>of</strong> the missing<br />

pottery was buried on site, a taphonomic regime resulting in substantial sherd attrition was<br />

inferred, leading to the conclusion that there was high potential for taphonomic bias in the<br />

recovered sample. Also, accumulations-based inferences that use quantities <strong>of</strong> recovered<br />

pottery to estimate intensity/duration <strong>of</strong> occupation are not applicable to data <strong>of</strong> this<br />

quality. This finding led to a question being raised in regard to Wickler’s inference <strong>of</strong> low<br />

intensity <strong>of</strong> occupation for similar sites in the Buka region. Also, comparisons using<br />

relative abundances <strong>of</strong> styles should be made with caution, particularly when comparing<br />

with well-preserved sites such as the Arawe Islands stilt villages and the Mussau stilt<br />

villages, where preservation appears to be extremely good and taphonomic bias is less<br />

likely. Similarly, comparison with terrestrial sites should be made with extreme caution,<br />

as while biases are likely in heavily gardened sites like those <strong>of</strong> the Reef/Santa-Cruz area,<br />

we do not yet know whether the biases are the same, or whether they are likely to create<br />

differences in stylistic composition <strong>of</strong> assemblages.<br />

How fragile is the site type? Chapter 5 informed on the state <strong>of</strong> preservation <strong>of</strong><br />

samples in a particular setting, the Roviana lagoon, where wave exposure is uniformly low<br />

and sea level history is similar all sites in the study. Even in this sheltered setting, the sites<br />

are in a very poor state <strong>of</strong> preservation, although visibility for those that remain exposed<br />

on the surface is high. This finding is <strong>of</strong> fundamental importance to the central question<br />

<strong>of</strong> this thesis, how to interpret the Lapita gap in Near-Oceania Solomon<br />

Islands/Bougainville. If the sites are poorly preserved in this sheltered setting, along most<br />

other coastlines exposure to a height level in relation to wave processes where<br />

archaeological visibility is high would result in total destruction <strong>of</strong> the ceramic component<br />

in most cases. Along such coastlines, where sea levels have fallen since Lapita times we<br />

489


should be looking for vestiges <strong>of</strong> sites such as stormwash accumulations rather than in-situ<br />

deposits, or deeply submerged sites if there is a history <strong>of</strong> relative rise in sea levels.<br />

<strong>The</strong> low state <strong>of</strong> completeness <strong>of</strong> the Roviana intertidal pottery provides some<br />

assurance that sherd counts are mostly independent observations. Sherd counts thus<br />

provide a useful measure <strong>of</strong> observation sample size, and the relative frequencies <strong>of</strong> pottery<br />

attributes in site samples. It should be noted that the same cannot be said <strong>of</strong> well preserved<br />

sites, where many sherds in the sample can derive from a single vessel.<br />

Chapter 6 extended the analysis <strong>of</strong> site preservation, adding the wave exposure<br />

component to a model <strong>of</strong> preservation. While information on prevailing winds conflicted,<br />

and environmental information on the extent and depths <strong>of</strong> reefs, sheltering sediments and<br />

seagrass beds was not <strong>of</strong> sufficient quality to generate precise comparative results, it is<br />

clear that none <strong>of</strong> the sites where pottery was found in quantity have more than 6 km <strong>of</strong><br />

wave exposure, and most have mitigating factors that prevent any substantial waves from<br />

affecting the sites at low tide. Thus the results <strong>of</strong> Chapter 5 can be extended to other<br />

regions as follows: where fetch is greater than 6km, and sea-levels have undergone a<br />

similar fall as in the Roviana setting, preserved intertidal ceramic sites are not expected to<br />

have survived except as lithic scatters or storm-redeposited material. This is just a<br />

hypothesis at present given some <strong>of</strong> the measurement uncertainties, but provides a rule-<strong>of</strong>-<br />

thumb starting point from which future survey <strong>of</strong> this sort can be targeted towards<br />

relatively sheltered locations.<br />

Analysis <strong>of</strong> formation process evidence in Chapter 7 considered a series <strong>of</strong> models<br />

<strong>of</strong> cultural formation processes, followed by a consideration <strong>of</strong> the evidence pertaining to<br />

natural or taphonomic processes as they affected ceramics. <strong>The</strong> absence <strong>of</strong> adjacent<br />

evidence on land, and the absence <strong>of</strong> erosion features other than solution scarring <strong>of</strong> Plio-<br />

Pleistocene upraised reefs are evidence against erosion <strong>of</strong> terrestrial sites into the sea. <strong>The</strong><br />

fresh state <strong>of</strong> preservation <strong>of</strong> Acropora corals on the upraised shore platform to landward<br />

<strong>of</strong> most sites suggests a recent high stand, the upper limit <strong>of</strong> coral growth, at which time<br />

490


the site locations would have been s<strong>of</strong>t backreef lagoonal sediments unaffected by swash<br />

processes. Paleoshoreline data (Mann et al. 1998) strongly support this interpretation, and<br />

suggest a high stand <strong>of</strong> at least 1.5m above present at 3000 BP, probably more.<br />

Subsequent sea-level fall led to to emergence <strong>of</strong> the backreef ceramic sites into the swash<br />

zone, such as it is. Coral growth accretions on some sherds also support this conclusion,<br />

as no modern coral growth occurs at these levels. Additional evidence against terrestrial<br />

settlement is complete absence <strong>of</strong> acid-leaching <strong>of</strong> carbonate grains (disregarding surface<br />

etching caused by cleaning sherds in acetic acid).<br />

Taking all these factors together, including the recent information from geology<br />

which was unavailable at the outset <strong>of</strong> the study, coral artificial islets seem an unlikely<br />

formation process for these sites, as the water would have been at least chest-deep at low<br />

tide, and probably deeper, requiring phenomenal labour to construct walling for a<br />

settlement. Stilt dwellings over shallow water, with piles set in s<strong>of</strong>t sediments now<br />

removed by swash processes as a result <strong>of</strong> uplift, are the most likely scenario.<br />

<strong>The</strong> artefact inventory <strong>of</strong> all intertidal sites was similar, dominated by pottery and<br />

lithic manuports, with rare adzes, flakes <strong>of</strong> meta-basalt or metamorphosed mudstone, very<br />

rare chert flakes, and a variety <strong>of</strong> abrader manuports. Coral abraders were not seen, but<br />

the corraline gravels (dominated by small Acropora fragments)forming the matrix <strong>of</strong> most<br />

sites would have made these difficult to find. <strong>The</strong>re was no evidence for site abandonment<br />

in the artefact inventory, or none that had survived postdepositional processes. All<br />

complete adzes were heavily resharpened in comparison to one unused preform in a private<br />

collection, suggesting sites were the focus <strong>of</strong> extended settlement resulting in the discard<br />

<strong>of</strong> ceramics, ovenstones/net weights and worn out adzes.<br />

Taphonomic analysis and spatial analysis (the latter at Zangana) <strong>of</strong>fered support<br />

for a transportation model <strong>of</strong> sherd attrition, while presence <strong>of</strong> ferromagnesian mineral<br />

grains in a sediment sample from the coralline Hoghoi shoreline suggested dissolution<br />

was also a factor in sherd attrition at some stage in the history <strong>of</strong> the sites. Taphonomic<br />

491


analysis <strong>of</strong> sherd samples most clearly supported the collector and collection intensity<br />

models, suggesting the potsherd scatters are highly sensitive to collection events. This<br />

finding is important for archaeological resource management and also places an additional<br />

burden on future investigators to thoroughly document the provenance <strong>of</strong> each item<br />

collected, and attend to desalination procedures. We cannot pick up a sample expecting<br />

to be able to go back later and get more <strong>of</strong> the same. This also suggests proactive<br />

education on the significance and fragility <strong>of</strong> the resource is desirable at the local level.<br />

Were artifact collectors to be successful in illegally selling items collected from such sites,<br />

we could expect the accessible surface-site information resource to rapidly degrade to an<br />

irretrievable level.<br />

Taphonomic analysis in Chapter 8 also provided information on the types <strong>of</strong> bias<br />

likely to be present in the ceramic samples. Strength variation between temper classes as<br />

evidenced by sherd size was low, but sherd thickness was clearly related to sherd size,<br />

suggesting thinner sherds are weaker and break more easily, and that thinner pottery is<br />

likely, or thinner parts <strong>of</strong> pots are likely to be underrespresented in the lag deposits as a<br />

result. Data was presented showing that some forms were stronger than others: Miho style<br />

pottery tended to be thicker at the neck and thin and fragile at the lip, suggesting this style<br />

is principally represented by neck sherds. Gharanga-style vessels with short, heavily-<br />

excurvate-to-rolled rims preserve well due to form strength, and lips <strong>of</strong> this form are thus<br />

more commonly attached to rims, leading to a larger count <strong>of</strong> Lip/Rim/Neck/Shoulder<br />

sherds.<br />

<strong>The</strong> analysis <strong>of</strong> pottery form variability and function in Chapter 8 focused on<br />

vertical contour <strong>of</strong> the upper vessel, and also used negative evidence, the absence <strong>of</strong> flat<br />

based sherds and stands (although there was one possible stand from Zangana). Six broad<br />

functional vessel form classes were identified on the basis <strong>of</strong> upper body contour<br />

variation, and a seventh was defined using body sherd thickness and form. <strong>The</strong>se classes<br />

were not all mutually exclusive, as some <strong>of</strong> the larger and more robust upper vessel sherds<br />

492


perhaps should have been classed together with the Form 7 sherds. <strong>The</strong> aims <strong>of</strong> the<br />

functional classification were the identification <strong>of</strong> a general purpose class that can be<br />

expected to have high use rates, exposure to thermal shock and short use life/ high discard<br />

rate, for use in form-controlled seriation analysis. A sub-class <strong>of</strong> Form 2 and a sub-class<br />

<strong>of</strong> Form 6 were judged to best meet these criteria, and were used for seriation analysis in<br />

Chapter 12.<br />

In Chapter 9 decorative variability was explored, and the Zangana site was split<br />

into two areas on the basis <strong>of</strong> decorative differences. Class 1 linear motifs<br />

(bounded/defined with double-line zone-markers, either dentate or incised) were largely<br />

restricted to the Honiavasa site, where they always occurred in association with carinated<br />

Form 1 vessels, as did a band <strong>of</strong> circular nubbins at the neck. It was these sherds that had<br />

led to the assignment <strong>of</strong> the site to the Lapita period from first discovery. Also present in<br />

some quantity were everted-rim necked vessels with either lip notching/impression <strong>of</strong><br />

various sorts, or a band <strong>of</strong> pinching on the outer edge <strong>of</strong> the lip.<br />

<strong>The</strong>re were only two examples <strong>of</strong> deformation <strong>of</strong> the lip into a wave form at<br />

Honiavasa, both <strong>of</strong> these coarsely executed, whereas this decorative technique was<br />

ubiquitous at Nusa Roviana, Paniavile and Zangana. Deformation <strong>of</strong> the lip into a<br />

discontinuous series <strong>of</strong> small spout-like decorative finger impressions was most common<br />

in Miho site, and occurred in all other sites for which sample size was adequate. <strong>The</strong>re<br />

were only two occurrences <strong>of</strong> this attribute at Honiavasa, compared with 22 occurrences<br />

at Miho. Class 2 linear motifs on the shoulder or rim were common at Miho, Paniavile and<br />

Zangana South, and absent from Honiavasa. A single band <strong>of</strong> fingernail pinching at the<br />

neck was common in the same sites as Class 2 linear motifs, and <strong>of</strong>ten occurred on the<br />

same sherds. This complex <strong>of</strong> decoration was termed the Miho decorative class or “Miho<br />

style” after the site where it was first recognised.<br />

<strong>The</strong> decorative attribute <strong>of</strong> a single punctate band at the neck was common in two<br />

sites, and was associated either with tall weakly everted rims above a slight neck<br />

493


estriction (Kopo style) or with short heavily everted rims (Gharanga style), the latter <strong>of</strong>ten<br />

associated also with multiple bands <strong>of</strong> fingernail pinching on the shoulder. Gharanga/Kopo<br />

style sherds showed a tendency towards thinner necks than Miho style, supporting the<br />

conclusions in Chapter 7 regarding the nature <strong>of</strong> stylistic-taphonomic bias.<br />

A heterogeneous sample <strong>of</strong> applied decoration proved difficult to classify and was<br />

omitted from the attributes used in seriations, although there is some discussion <strong>of</strong> possible<br />

temporal variability given below.<br />

Analysis <strong>of</strong> lip impression/crenation/notching in relation to lip form found a<br />

correlation between lip orientation angle and the location <strong>of</strong> decoration which cross-cut<br />

drastic differences in vessel form and decoration, suggesting that lip impression attributes<br />

could cause faulty ordering <strong>of</strong> a seriation. Location <strong>of</strong> lip impression bands seemed to be<br />

correlated with lip orientation, but as this measurement was dependent on rim orientation<br />

measurements, which had poor precision, this requires confirmation from an enlarged<br />

sample <strong>of</strong> sherds <strong>of</strong> 10% EVE or greater (below which rim orientation is difficult to<br />

measure).<br />

Neck decoration for all styles (Honiavasa, Miho and Gharanga-Kopo styles) was<br />

dominated by a band <strong>of</strong> circular elements encircling the neck, while technique <strong>of</strong><br />

execution varied between these styles. Within an evolutionary definition <strong>of</strong> style and<br />

function, following Dunnell, this suggests, if these are temporal types, that the pattern is<br />

constrained and remaining constant, while the technique <strong>of</strong> execution varies, either<br />

through drift or selection. <strong>The</strong> dominance <strong>of</strong> this pattern at the neck <strong>of</strong> the pot and the<br />

temporal constancy suggests the pattern is fundamental to the decorative ideas <strong>of</strong> the<br />

potters, and is functional in some way. A parallel was seen with Spriggs’ “changing face<br />

<strong>of</strong> Lapita” in that these circular elements could be seen as the eyes in the Lapita face, and<br />

a protective ide<strong>of</strong>unction was suggested, where the eyes were an example <strong>of</strong> the<br />

“technology <strong>of</strong> enchantment or the enchantment <strong>of</strong> technology” (Gell 1992), perhaps<br />

conferring protection from illness or poisoning to the contents <strong>of</strong> the pot and to eaters <strong>of</strong><br />

494


the contents.<br />

Analysis <strong>of</strong> lithics in Chapter 10 focussed on petrographic classification <strong>of</strong> lithic<br />

manuports, analysis <strong>of</strong> size distribution <strong>of</strong> these, morphological and petrographic<br />

description <strong>of</strong> adzes, and petrographic description <strong>of</strong> sandstone abrader fragments. <strong>The</strong><br />

diversity <strong>of</strong> sources represented suggests enormous potential <strong>of</strong> such sites for lithic<br />

sourcing studies, as a means <strong>of</strong> reconstructing patterns <strong>of</strong> raw material transport in the<br />

past. This conclusion should add archaeological value to these intertidal/underwater<br />

surface scatters, and encourage more intensive recording and sampling in the course <strong>of</strong><br />

future research and resource management.<br />

Some <strong>of</strong> the unfractured water-rounded lithic manuports seemed too small to have<br />

been procured as ovenstones, although small stones are sometimes collected for this<br />

purpose in Oceania today (Nojima 2002: Pers. Comm.). Another functional possibility is<br />

as net weights for large nets such as used ethnographically for turtle or dugong, both <strong>of</strong><br />

which are common today in the Lagoon, dugong especially in extensive seagrass meadows<br />

adjacent to Hoghoi.<br />

While most manuports probably were sourced from Rendova, some may have<br />

originated from further afield (Viru Harbour or Kolombangara, both about 50km distant),<br />

although this suggestion requires testing through examination <strong>of</strong> likely source streams.<br />

Adzes are from a diversity <strong>of</strong> sources, showing a preference for metamorphosed<br />

siltstones or metamorphosed fine-grained volcanic rocks. While some <strong>of</strong> these adze rocks<br />

are conceivably from Tertiary sedimentary formations <strong>of</strong> the forearc chain <strong>of</strong><br />

Ranonnga/Southern Rendova/Tetepare, 15km or more distant from Roviana, the<br />

recrystallization evident petrographically may indicate an older origin further afield,<br />

perhaps among the complex geology <strong>of</strong> Choiseul, or possibly some <strong>of</strong> these result from<br />

more local contact metamorphism associated with plutonic intrusions.<br />

Adze forms are consistent with the Buka/Nissan Lapita reef-site evidence, and<br />

suggest heritable continuity with Lapita, supporting Reeve’s contention that the Paniavile<br />

495


material was Lapita-derived. This evidence for heritable continuity was also useful for<br />

seriation, as a demonstration <strong>of</strong> heritable continuity is important for evolution-based<br />

techniques such as seriation. <strong>The</strong> Near-Oceania stone adze sample from the Lapita period<br />

is now dominated by the materials obtained from reef/intertidal sites, suggesting surface<br />

archaeology holds the key to acquiring a large sample with which to study variability <strong>of</strong><br />

this rare artefact class in detail.<br />

Petrography <strong>of</strong> abrasives indicated procurement from a variety <strong>of</strong> sources, and the<br />

diversity <strong>of</strong> sources (no repeats) suggests the parent population <strong>of</strong> sources utilized is large.<br />

Our sample <strong>of</strong> these sorts <strong>of</strong> materials is far from saturated. Detailed petrographic and<br />

morphological comparisons with materials recovered by Wickler from Buka would be a<br />

rewarding exercise.<br />

Analysis <strong>of</strong> spatial structure (Chapter 11) in the ceramic data and the lithic Hoghoi<br />

data identified evidence for size and density sorting, supporting the sherd transport natural<br />

formation process model <strong>of</strong> Chapter 7. <strong>The</strong> analysis also looked for structured distribution<br />

<strong>of</strong> tempers and styles. <strong>The</strong>re was a clear separation <strong>of</strong> Miho decorative class from<br />

Gharanga/Kopo decorative class at Zangana, supporting division into Zangana North and<br />

Zangana South for seriation. By contrast, wave-deformation <strong>of</strong> the lip occurred ten times<br />

in each <strong>of</strong> the two halves, suggesting this attribute had a different temporal trajectory to<br />

the Miho decorative class despite association on some sherds with unbounded incised rim<br />

decoration.<br />

<strong>The</strong>re was no strong evidence for a spatial separation <strong>of</strong> the exotic quartz-calcite<br />

temper class from other tempers, which suggests, in conjunction with TL dating results,<br />

stylistic analysis, and occurrence in all sites, that this class <strong>of</strong> temper had an extended<br />

period <strong>of</strong> production and importation, despite the low sherd count.<br />

In Chapter 12 three lines <strong>of</strong> chronological evidence were integrated: these being<br />

radiocarbon dates; thermoluminescence; and ceramic seriation. C 14 data were interpreted<br />

as allowing reasonable confidence that the Roviana intertidal sites have an occupation<br />

496


span <strong>of</strong> at least 160 years, unless there is an old-wood problem with the determination on<br />

smoke-derived carbon from the surface <strong>of</strong> a sherd in the Hoghoi site. Direct dating was<br />

successful in two out <strong>of</strong> three attempts, with one case <strong>of</strong> dating <strong>of</strong> a sub-fossil organic<br />

inclusion clearly documented. <strong>The</strong> prospects for AMS direct dating <strong>of</strong> pottery are good,<br />

and throw the spotlight onto developing techniques for locating additional samples within<br />

sherds, both from the current collection and from future samples. Demonstrating that it is<br />

now possible to derive independent chronological data from the physical properties <strong>of</strong><br />

sherds and lithics in Near Oceania in this way, using TL and Radiocarbon, places added<br />

archaeological value on “disturbed” archaeological contexts such as intertidal lag scatters.<br />

<strong>The</strong> addition <strong>of</strong> eight TL dates expands the minimum occupation span <strong>of</strong> the<br />

intertidal pattern to 611years at 2 s.d.. If the TL dates from the Quartz-calcite temper are<br />

read as relative ages, disregarding corrections for anomalous fading, these suggest<br />

Honiavasa is the oldest site, while Miho-style pottery from Miho, Zangana and Paniavile<br />

is mostly slightly younger, with the youngest dates coming from Paniavile and Gharanga.<br />

More TL data are needed to test the notion that the uncorrected ages provide a relative<br />

chronology due to origin from a common source, and TL/carbon pairs would be an ideal<br />

scenario. If a systematic model <strong>of</strong> Luminescence fading can be developed for the<br />

feldspathic component <strong>of</strong> the quartz-calcite temper, then TL dating could become a key<br />

tool for defining Lapita chronology in the Western Solomon Islands.<br />

Seriation using CA produced different orderings depending on which attributes<br />

were included, and the results <strong>of</strong> seriation from which lip impressions were excluded were<br />

preferred, but this cannot be independently confirmed on current data. <strong>The</strong> tendency for<br />

ceramic variability to coalesce into three styles (Honiavasa, Miho and Gharanga/Kopo was<br />

partly an result <strong>of</strong> choice and sample sizes <strong>of</strong> attributes, but is likely also to be a<br />

consequence <strong>of</strong> a historically incomplete sample. Honiavasa site has very little in common<br />

ceramically with the other sites, although enough to suggest heritable continuity. Taken<br />

together with continuity in settlement types, the preferred explanation for the three styles<br />

497


is that we are missing the transition from Honiavasa utilitarian ceramics to Miho and<br />

Gharanga/Kopo utilitarian ceramics, unless some <strong>of</strong> the ubiquitous attributes span this gap,<br />

for example lip deformation on plain rims. Additional sampling <strong>of</strong> the regional record is<br />

required, including areas beyond the current sampling frame, as concluded on other<br />

evidence in Chapter 3 also. Should these stylistic discontinuities persist with additional<br />

sampling, this conclusion would have to be reevaluated.<br />

When the three different chronologies are synthesised, a Honiavasa>Miho-<br />

style>Gharanga/Kopo style sequence seems most likely. Wave-deformation <strong>of</strong> the lip<br />

seems likely to begin during the production span <strong>of</strong> the Miho style and continue till tall-rim<br />

Kopo-style production commences. Gharanga-style short-rim vessels may be<br />

contemporaneous functional variants <strong>of</strong> the Gharanga/Kopo period, or may be a<br />

development out <strong>of</strong> Kopo style, with multi-band pinching <strong>of</strong> the shoulder, and thinner<br />

construction being innovations following the reduction <strong>of</strong> rim height.<br />

Under Balfet’s model as applied by Irwin (Balfet 1965, Irwin 1985), this sequence<br />

may document replacement <strong>of</strong> domestic idiosyncratic low-skill production with specialist<br />

standardised technically-accomplished production. <strong>The</strong> chunky Form 1 carinated jars in the<br />

Honiavasa site with individualistic decoration, and heavy Form 6 pots, have been replaced,<br />

if this series is correct, by progressively thinner and simpler, more standardised pots <strong>of</strong> a<br />

more general-purpose nature, with better thermal properties as a result <strong>of</strong> reduction in<br />

thickness, also potentially allowing more rapid drying prior to firing.<br />

External Comparisons:<br />

Having critiqued the use <strong>of</strong> presence-absence data for comparing subsamples <strong>of</strong> past<br />

behaviour in Chapter 1 and 2, I am unwilling to make such comparisons using the<br />

Honiavasa data as I believe these will simply mislead. <strong>The</strong> Honiavasa sample contains a<br />

498


small corpus <strong>of</strong> carinated vessels with Lapita motifs, but sample size is insufficient for<br />

relative-frequency based motif comparisons also. Comparisons are thus restricted to some<br />

general statements. <strong>The</strong> Honiavasa carinated vessels with Class 1 motifs have clear<br />

homologous similarity to the Lapita ceramic horizon. <strong>The</strong>re are no flat bases or cylinder<br />

stands in the sample, but this does not mean these were absent in the past. <strong>The</strong> complex<br />

curvilinear motifs characteristic <strong>of</strong> early Lapita according to Best are not present in the<br />

sample. Ishimura’s recent analysis, by contrast, has these techniques <strong>of</strong> decoration evolving<br />

over a more lengthy period across the temporal dimension <strong>of</strong> the Lapita horizon.<br />

A number <strong>of</strong> Anson’s motifs, or similar motifs are present in the small Honiavasa<br />

carinated vessel sample. Mead motifs 16, something similar to motif 11 or 16.3, something<br />

similar to motif 15, and a geometric motif similar to 17.1 or 30 are present among the small<br />

sample <strong>of</strong> carinated Form 1 vessels. Mead’s 3-dimensional design elements N1.1, VB2,<br />

TB3.1, and TB3.3 occur in the Honiavasa sample. Design elements 1.2, 5 and 6 occur at<br />

Honiavasa. <strong>The</strong>se characteristics also serve to link Honiavasa to the Lapita ceramic<br />

horizon.<br />

Carinated vessel forms at Honiavasa are characteristic <strong>of</strong> Summerhayes’ early<br />

Lapita in the Bismarck Archipelago (Summerhayes Form V), but the rare occurrences <strong>of</strong><br />

dentate-stamping argue for a later date under the Summerhayes scheme. While this site<br />

was initially described as late-Lapita (Felgate 2002) I would now prefer to have a detailed<br />

regional sequence as a basis for cross matching, and am less confident in the temporal<br />

constructs <strong>of</strong> others than at the outset <strong>of</strong> analysis.<br />

I initially thought that the Roviana intertidal ceramics must date to around 500BC<br />

or younger, based on the information from Watom and Buka, and the first C14 date from<br />

Paniavile was consistent with this assignment. Feathers, however, was forced to reject his<br />

measurements from Honiavasa as too old, when he compared the results obtained to the<br />

expected age. Since then, a date likely to be around 800BC from Hoghoi has suggested<br />

that the Honiavasa Lapita subtantially predates 500BC, also that the TL measurements<br />

499


were giving a reasonable ordination chronology, and that Lapita is effectively over by<br />

about 800BC at Roviana Lagoon.<br />

As a result <strong>of</strong> this information, combined with the results <strong>of</strong> a critical review <strong>of</strong> the<br />

underpinnings <strong>of</strong> current constructs <strong>of</strong> Lapita variability, I am now inclined to put the<br />

absolute age <strong>of</strong> the Honiavasa carinated pottery at no younger than 850BC, and possibly<br />

older. Miho style pottery may be being produced by about 800BC, and my inclination,<br />

although as discussed in Chapter 12 this is unconfirmed by solid evidence, is to have<br />

Gharanga/Kopo style produced sometime between 800BC and 0.AD.<br />

Several occurrences <strong>of</strong> compound rims at Honiavasa provide similarity with rim<br />

pr<strong>of</strong>iles from Buka (especially Wickler 2001: Figure 4.2k from site DAF). <strong>The</strong>se rims are<br />

listed by Sand as characteristic <strong>of</strong> the Southern Lapita Province (Sand 2000), and the Buka<br />

and Roviana materials provide a suggestion that they are a more widespread trait. Poulsen<br />

calls these “flange” rims for Tonga, and these are Mead’s 3-dimensional design element<br />

2.8, defined from a Sigatoka sample, so they are clearly widely distributed, although the<br />

best examples are from the large whole vessels recently found in New Caledonia. Given<br />

the widespread distribution <strong>of</strong> these I hesitate to attribute them to a late date, and regard<br />

them as potentially part <strong>of</strong> the repertoire <strong>of</strong> rim/vessel forms <strong>of</strong> the initial spread into<br />

Remote Oceania, and thus having an origin <strong>of</strong> similar antiquity to the curvilinear/face<br />

motifs such as Mead M33. Alternatively, the widespread occurrence <strong>of</strong> these rim forms<br />

may arise through reticulate processes in RemoteOceania and the Near-Oceanic Solomon<br />

Islands/Buka. I am tempted to call these an indication <strong>of</strong> a “Central Lapita Province”,<br />

rather than Western Lapita, the latter something <strong>of</strong> a misnomer when Western Lapita is<br />

found as far east as Tonga (eg. Burley & Dickinson 2001).<br />

It is possible that the Roviana early ceramic sequence, in spite <strong>of</strong> the low number<br />

<strong>of</strong> sites, few dates and patchy sample, is amplifying the temporal resolution obtainable<br />

from excavated samples from elsewhere in Near Oceania for the Lapita/post-Lapita<br />

transition, as predicted by the extended review in the opening <strong>chapter</strong>s <strong>of</strong> the nature and<br />

500


potential <strong>of</strong> seriation <strong>of</strong> surface sites as an approach to the archaeology <strong>of</strong> a region. <strong>The</strong><br />

Garua pottery from FSZ looks like a mixture <strong>of</strong> attributes that are spatially separated in the<br />

Roviana surface sites. Wavy stamping, as illustrated by Summerhayes for the Garua sample<br />

(Summerhayes 2000a:146, Figure 9.2, sherds 1620, 1101, 1927) occurs on several<br />

Honiavasa sherds, in similarly abstract arrangements, with hints <strong>of</strong> use as a latitudinal zone<br />

marker in some cases. Also, Summerhayes' Sherd 1275, for instance, is a typical Roviana<br />

wave-deformed lip, occurring in high frequency at Miho, Paniavile, Zangana and Nusa<br />

Roviana, but only rarely and in an odd, coarse, heavy variant at the Honiavasa Lapita site.<br />

<strong>The</strong> separation <strong>of</strong> Honiavasa and Paniavile using linear motifs provides another clue<br />

that in the Roviana surface-collected data we may be seeing an amplification <strong>of</strong> temporal<br />

resolution in the relatively unrestricted landscape with plenty <strong>of</strong> places to move a stilt<br />

settlement to. Miho style is dominated by variants <strong>of</strong> Mead motif 18, while motif 24 also<br />

occurs several times. On the shoulder <strong>of</strong> Miho-style vessels it is not uncommon to see a<br />

simplified form <strong>of</strong> motif 16, being just the diagonal lines forming a zigzag, expressed using<br />

the single-line (Mead GZ3) rather than twinned lines. <strong>The</strong> motif listed as Mead M16.1<br />

(Green 1979) (different to that illustrated by Mead) is present at Paniavile, where Miho<br />

style predominates, and is associated with pinching at the neck, a feature diagnostic <strong>of</strong> the<br />

Miho style.<br />

This idea <strong>of</strong> temporal amplification obtains additional support from the<br />

occurrences <strong>of</strong> Gharanga/Kopo style punctation elsewhere. <strong>The</strong> characteristic punctation<br />

or stick impression diagnostic <strong>of</strong> Gharanga/Kopo style, where the still-plastic clay is<br />

deformed inwards slightly as a bump on the interior surface <strong>of</strong> the neck region occurs on<br />

more than 60 sherds in the sample, and is completely absent from the large Honiavasa<br />

vessel sample. This decoration is illustrated by Anson from Ambitle (Anson 1983: Figure<br />

X), and occurs on two sherds at site DES on Nissan (Wickler 2001:120), and at the EKQ<br />

site on Mussau (Kirch et al. 1991: Figure 4c). Compared to these samples, the Roviana<br />

501


sample, particularly from Gharanga and Hoghoi, represents the parent lode, unless this is<br />

simply analogous similarity, but if the latter is the case, why is it so rare elsewhere?<br />

<strong>The</strong> single quartz-calcite tempered example <strong>of</strong> this class <strong>of</strong> decoration may indicate<br />

an extended zone <strong>of</strong> production, design emulation, or exchange <strong>of</strong> potters as well as pots,<br />

as suggested by Summerhayes also from other data. Most <strong>of</strong> these vessels have placered<br />

volcanic temper, and it is not certain that there is much benefit in petrographic comparisons<br />

with the other examples from Buka and Papua New Guinea, although it never hurts to<br />

look.<br />

Gharanga-style multiple bands <strong>of</strong> fingernail impression on the shoulder occur in low<br />

frequency in many samples elsewhere, and has generally been investigated at the level <strong>of</strong><br />

decorative technique rather than at the level <strong>of</strong> structured pattern across the vessel (e.g.<br />

Anson 1983:Figure XI) although Bedford is much more specific in this respect (see for<br />

example Bedford 2000: early and late Ifo ware, Figures 7.2-7.4). Until chronologies are<br />

more clearly defined it will be difficult to distinguish between homologies and analogies<br />

in respect <strong>of</strong> these simple decorative techniques, and Bedford’s reassessment <strong>of</strong> the incised<br />

and applied relief “tradition” is apposite here too, providing a caution against jumping to<br />

hastily to conclusions and making unwarranted connections.<br />

If there is any connection, and/or if decorative technique and pattern is separating<br />

out by site at Roviana, where it is mixed in with other pottery to the northwest or Vanuatu,<br />

either the Roviana surface scatters are amplifying chrono-stylistic structure in the data and<br />

thus temporal resolution, where this is more temporally mixed in excavated samples to the<br />

northwest, or the trajectories <strong>of</strong> ceramic change are so different between Roviana and the<br />

Buka/New Britain area that Roviana pottery may have changed beyond recognition as<br />

Lapita by 800BC, while Lapita-style is still supposedly thriving in the Bismark<br />

Archipelago.<br />

Whether Gharanga/Kopo pottery is thus a valuable index fossil (O'Brien & Lyman<br />

2000), and similar sites further north are yet to be discovered, or whether this is just<br />

502


analogous similarity (unlikely in view <strong>of</strong> the rarity <strong>of</strong> occurrence further north) cannot be<br />

resolved until temporal variability is better understood at the local level in each <strong>of</strong> these<br />

areas. <strong>The</strong> preferred explanation is that we have yet to fully sample and define temporal<br />

variability fully for the Lapita period and what came after, in New Britain, at Buka, and at<br />

Roviana Lagoon, and that the current samples from throughout the region, including<br />

Roviana and Buka and the Bismarck Archipelago, are insufficiently historically complete<br />

and site-structured to allow the sort <strong>of</strong> high-resolution chronology that can be used to<br />

discriminate between homologous and analogous similarity <strong>of</strong> simpler designs between<br />

regions.<br />

Late prehistoric adzes from Gharanga and Paniavile tie into Specht and Wickler’s<br />

terrestrial chronology, and suggest intertidal settlement may have continued until circa<br />

1500 AD, but this is speculative as the adzes are reported to have come from the intertidal,<br />

but are not securely provenanced. Even if these were found in the intertidal scatters, this<br />

does not mean they are the same age as the bulk <strong>of</strong> pottery in those sites, as surface<br />

collections, to a greater extent than some buried sites can easily incorporate materials from<br />

various periods.<br />

<strong>The</strong> Lapita Gap as an Area <strong>of</strong> Low Probability <strong>of</strong> Detection :<br />

<strong>The</strong> Roviana study shows that Lapita is most likely continuously distributed across near<br />

Oceania, because the gap in the recorded distribution can now be seen to reflect<br />

preservation and visibility rather than past behaviour. Now that we know the poor state<br />

<strong>of</strong> preservation <strong>of</strong> the emergent Roviana stilt villages, and the likely depth <strong>of</strong> water in<br />

which they were constructed, the absence <strong>of</strong> evidence from Choiseul, Ysabel and Ngela<br />

is explicable by the geological evidence for subsidence in those areas (Bruns et al. 1986).<br />

If the main chain adjacent to the New Georgia group is subsiding while Roviana is<br />

uplifting, sites at the sorts <strong>of</strong> locations evident in the Roviana data will be either grown<br />

503


over with coral or buried in s<strong>of</strong>t reefal detrital sediments, and thus be both well preserved<br />

and invisible, like the Mulifunua Ferry Berth site in Western Samoa, discovered by<br />

fortuitous dredging. In areas with more substantial uplift than Roviana, such as the forearc<br />

chain <strong>of</strong> Rendova, Tetepare and Ranonnga, preservation is likely to be worse than in the<br />

Roviana Lagoon, in some areas through wave exposure, but also because uplift has<br />

brought the pottery through the swash zone where it cannot survive wave exposure. This<br />

can be expected to have substantially reduced probability <strong>of</strong> detection. Evidence suggesting<br />

a Lapita gap in the Near-Oceanic Solomon Islands as avoidance or leap-frog colonization<br />

is thus substantially eliminated.<br />

This thesis can be tested by looking in places where high probability <strong>of</strong> past<br />

settlement, archaeological visibility, preservation and survey methods combine to promote<br />

a high probability <strong>of</strong> site detection. <strong>The</strong> scale <strong>of</strong> survey required is suggested in Chapter<br />

3 to be <strong>of</strong> the order <strong>of</strong> 900km 2 <strong>of</strong> lagoon, or 40 reef passages, or about 150km <strong>of</strong> coastline<br />

with high detectability, to detect a sample <strong>of</strong> ten or more sites with good samples <strong>of</strong><br />

Lapita-era and a larger number <strong>of</strong> derivatives <strong>of</strong> Lapita. With this sort <strong>of</strong> information in<br />

hand, the behavioural questions that require us to be able to tell time with high temporal<br />

resolution (Spriggs 2001) including consideration <strong>of</strong> spatial factors such as site density and<br />

mobility models (Anderson 2002) can be matched to appropriate data.<br />

Intertidal-Zone and Shallow-Water Archaeology:<br />

This thesis has been as much about figuring out how to proceed in researching an intertidal<br />

and shallow water archaeological distribution as about the distribution <strong>of</strong> Lapita pottery.<br />

<strong>The</strong> key <strong>chapter</strong>s in this regard are those on survey (Chapter 3) and formation processes<br />

(Chapters 5, 6, 7 and 11). <strong>The</strong>se could all have been vastly improved had the<br />

environmental variables used (wave-exposure; sediment characteristics and vegetation<br />

cover; relative sea level and tidal fluctuations; and collection site micro-topography) and<br />

504


the locations <strong>of</strong> artefacts and manuports been recorded more systematically and in more<br />

detail. This would have necessitated either a lot more survey time or much better<br />

equipment (probably both). In retrospect, if accurate electronic survey equipment was<br />

being used to capture these environmental details, point-provenancing <strong>of</strong> all sherds and<br />

lithics would have been a simple matter too.<br />

<strong>The</strong> wet tropical environment <strong>of</strong> the Roviana study is hard on such gear, which<br />

must be carefully maintained to function well, may fail unexpectedly, and adds to the<br />

expense and logistics <strong>of</strong> fieldwork Many <strong>of</strong> the suggestions for practice might seem trivial<br />

to those working in urban environments or in the rural hinterland <strong>of</strong> industrialised societies,<br />

where power supplies are usually available either from the main grid or from vehicle<br />

batteries, and where electronic equipment is available either through <strong>University</strong><br />

departments or Resource Management Agencies or hire companies, and is <strong>of</strong>ten used as<br />

a matter <strong>of</strong> course. Some <strong>of</strong> these suggestions might be very difficult to implement,<br />

however, by resource management agencies in non-industrialised nations or regions, like<br />

Solomon Islands, particularly where funding is meagre. <strong>The</strong> core information, though, that<br />

cannot be reconstructed after the event <strong>of</strong> picking up a sample, is where the piece came<br />

from, which is brought home to me by analysis <strong>of</strong> data which reflects my own poor efforts<br />

in this regard.<br />

A low-tech solutions to such recording can include detailed triangulation mapping<br />

<strong>of</strong> sample locations using measuring tapes from a permanently marked baseline, and is thus<br />

possible with simple equipment, although time consuming. Desalination <strong>of</strong> porous materials<br />

can be done in rainwater, as long as some method <strong>of</strong> labelling artefacts to retain their<br />

provenance information can be devised. Even cheaper, if this cannot be done, is to<br />

encourage protection <strong>of</strong> the site at the local level, where possible, and to leave the items<br />

where they are unless there is an immediate threat to their survival. Non-destructive in-situ<br />

recording <strong>of</strong> form and decoration for research purposes is always an option if there is<br />

sufficient skilled labour available. In some circumstances marine growth on sherds will<br />

505


obscure the details <strong>of</strong> decoration, in which case laboratory cleaning might be required.<br />

Broad fabric classifications could also be done in the field in many cases using a 10x hand<br />

lens, if a small corner <strong>of</strong> each sherd is snapped, with a type-series retained for petrographic<br />

analysis. <strong>The</strong>se suggestions for practice may not all turn out to be entirely practical, but<br />

are the range <strong>of</strong> possibilities I would experiment with in any future fieldwork <strong>of</strong> this sort.<br />

<strong>The</strong> focus on formation processes, which involved treating sherds and lithics as<br />

sedimentary particles, required descriptive systematics adapted to that purpose. Sherd size,<br />

in particular, was recorded in more detail than is the norm in Pacific archaeology, as was<br />

lithic manuport size and form. Point provenancing <strong>of</strong> artefacts, and recording <strong>of</strong> orientation<br />

could be useful developments <strong>of</strong> this approach.<br />

While the search for undisturbed buried villages continues in Lapita archaeology,<br />

there has been little explicit consideration <strong>of</strong> the problem <strong>of</strong> what to do with these when<br />

they are found (exceptions to this trend are noted in the introduction to Chapter 11).<br />

Relatively well-preserved stilt-house remains have been found in the Arawes and Mussau,<br />

and terrestrial buried landscapes have been reported from Manus and Vanuatu, but none<br />

<strong>of</strong> these have been extensively excavated, because to do so would take an enormous<br />

amount <strong>of</strong> time, money and skilled labour. Settlement pattern studies in the Reef-Santa<br />

Cruz islands and Fiji using both the surface and excavation record have been rare<br />

exceptions rather than the rule, enabled in part by shallow burial. Investigation <strong>of</strong> deeply-<br />

buried places has tended to become a search for layercake chronology, largely due to the<br />

difficulties <strong>of</strong> excavating anything other than a small fraction <strong>of</strong> the areas in a<br />

reconnaissance mode. <strong>The</strong>re is nothing intrinsically wrong with this, provided stringent<br />

attention is paid to identification <strong>of</strong> formation processes, and age-depth relationships are<br />

carefully constructed rather than assumed to be readable as time. <strong>The</strong>se relatively small<br />

holes will seldom be easily demonstrated to contain representative samples <strong>of</strong> variability<br />

for any <strong>of</strong> the layers, due to limited spatial sampling and the difficulty <strong>of</strong> quantifying<br />

506


samples, an area in which this thesis has sought to make some progress.<br />

Well preserved ceramic sites yield a lot <strong>of</strong> sherdage, which can foster complacence<br />

that the sample obtained from a small excavation is representative <strong>of</strong> the settlement as a<br />

whole, both across space and through time. <strong>The</strong> analysis in Chapter 5 provides an example<br />

showing how distantly related physical quantity <strong>of</strong> ceramics and sample size can be. It is<br />

this aspect <strong>of</strong> a tendency to search for layer upon layer <strong>of</strong> buried villages that is the focus<br />

<strong>of</strong> much <strong>of</strong> the critique in Chapter 2, and the argument is put forward in various places in<br />

this thesis that the temporal resolution <strong>of</strong> a seriation chronology constructed from a large<br />

number <strong>of</strong> large samples will almost inevitably surpass that obtained from reconnaissance<br />

excavation squares.<br />

In making external comparisons using the Roviana data above, the suggestion is<br />

put forward that some <strong>of</strong> the difficulty in matching the Roviana series to that from other<br />

areas may be to do with differences in temporal resolution, where horizontal structure is<br />

separating out styles that are mixed in other buried contexts. An argument for large area<br />

excavations and surface collection has been championed by Green for many years, and in<br />

the lithics <strong>chapter</strong> (Chapter 10) it is clear that the stone adze sample from Lapita-era sites<br />

in Near Oceania is dominated by materials from surface collections (this probably holds<br />

true for remote Oceania also, if large-area excavations are counted as equivalent in the<br />

scope <strong>of</strong> sampling to surface collections).<br />

I put it to research archaeologists interested in Lapita that this is probably true also<br />

<strong>of</strong> pottery, it is just that we have been lax about quantifying sample size, as few have been<br />

willing to battle with the problem <strong>of</strong> establishing vessel brokenness and completeness, and<br />

to develop methods <strong>of</strong> assessing sample size in various sampling contexts using these<br />

properties <strong>of</strong> samples. For these reasons we should value the surface record, including lag<br />

deposits in the sea, especially in Near Oceania (where settlement in such locations in the<br />

past is now well documented), as a source <strong>of</strong> large, easily acquired samples <strong>of</strong> ceramics<br />

and lithics, that can potentially form the backbone <strong>of</strong> high-resolution chronologies on<br />

507


which so much behavioural inference depends.<br />

Landscapes which <strong>of</strong>fer surface exposure <strong>of</strong> samples <strong>of</strong> Lapita-age materials<br />

sufficiently large to allow exploration <strong>of</strong> ceramic variability, <strong>of</strong> which the New Georgia<br />

region is one, should not be too hastily bypassed in the noble quest for Wobst’s mythical<br />

single point where all the region’s culture-historical complexes can be found superposed,<br />

separated by sterile deposits.<br />

<strong>The</strong> End<br />

508


REFERENCES<br />

Allen, J., 1984. Pots and Poor Princes: A Multidimensional Approach to the Role <strong>of</strong><br />

Pottery Trading in Coastal Papua. In S. E. Van Der Leeuw and A. C. Pritchard<br />

(eds) <strong>The</strong> Many Dimensions <strong>of</strong> Pottery: Ceramics in Archaeology and<br />

Anthropology. Amsterdam: Universiteit van Amsterdam,<br />

Allen, J., Gosden C., 1991. Report <strong>of</strong> the Lapita Homeland Project. Canberra:<br />

Department <strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific Studies, ANU.<br />

Allen, J., Gosden C., White J. P., 1989. Human Pleistocene Adaptations in the Tropical<br />

Island Pacific: Recent Evidence from New Ireland, a Greater Australian Outlier.<br />

Antiquity 63: 548-61<br />

Allen, J., Specht J., Ambrose W. R., Yen D. E., 1984. Lapita Homeland Project: Report<br />

<strong>of</strong> the 1984 Field Season Department <strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific<br />

Studies, ANU, Canberra<br />

Ambrose, W., 1991. Lapita or Not Lapita: <strong>The</strong> Case <strong>of</strong> the Manus Pots. In J. Allen and C.<br />

Gosden (eds) Report <strong>of</strong> the Lapita Homeland Project. Canberra, Australia:<br />

Department <strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian<br />

National <strong>University</strong>, pp.103-12.<br />

Ambrose, W., Gosden C., 1991. Investigations on Boduna Island. In J. Allen and C.<br />

Gosden (eds) Report <strong>of</strong> the Lapita Homeland Project. Canberra, Australia:<br />

Department <strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian<br />

National <strong>University</strong>, pp.182-8.<br />

Anderson, A., 2002. Mobility Models <strong>of</strong> Lapita Migration. In G. R. Clark, A. J. Anderson<br />

and T. Vunidilo (eds) <strong>The</strong> Archaeology <strong>of</strong> the Lapita Dispersal in Oceania:<br />

Papers from the 4th Lapita Conference, June 2000, Canberra, Australia.<br />

Canberra: Pandanus Press, Research School <strong>of</strong> Pacific and Asian studies, ANU,<br />

pp.15-23.<br />

509


Anson, D., 1983. Lapita Pottery <strong>of</strong> the Bismarck Archipelago and Its Affinities. PhD<br />

thesis. <strong>University</strong> <strong>of</strong> Sydney.<br />

---, 1986. Lapita Pottery <strong>of</strong> the Bismarck Archipelago and Its Affinities. Archaeology in<br />

Oceania 21: 157-63<br />

---, 1987. Reply to Kirch Et Al. Archaeology in Oceania 22: 127-8<br />

---, 1990. Aspiring to Paradise. In M. Spriggs (ed) Lapita Design, Form and<br />

Composition: Proceedings <strong>of</strong> the Lapita Design Workshop, Canberra, Australia,<br />

December 1988. Canberra Australia: Department <strong>of</strong> Prehistory, Research School<br />

<strong>of</strong> Pacific Studies, Australian National <strong>University</strong>, pp.53-8.<br />

---, 2000. Reber-Rakival Dentate-Stamped Motifs: Documentation and Comparative<br />

Implications. New Zealand Journal <strong>of</strong> Archaeology 20(1998): 119-35<br />

Anthes, R. A., 1982. Tropical Cyclones, <strong>The</strong>ir Evolution, Structure and Effects.<br />

Meteorological Monographs No. 41. American Meteorological Society.<br />

Arnold, D. E., 1985. Ceramic <strong>The</strong>ory and Cultural Process. Cambridge <strong>University</strong> Press.<br />

Arqueologicas, P. N. D. P., 1970. Brazilian Archaeology in 1968: And Interim Report on<br />

the National Program <strong>of</strong> Archaeological Research. American Antiquity 35 (1): 1-<br />

23<br />

Aswani, S., 1997. Customary Marine Tenure and Artisanal Fishing in the Roviana and<br />

Vonavona Lagoons, Solomon Islands: <strong>The</strong> Evolutionary Ecology <strong>of</strong> Marine<br />

Resource Utilization. PhD thesis. Anthropology, <strong>University</strong> <strong>of</strong> Hawaii.<br />

Balfet, H., 1965. Ethnographic Observations in North Africa and Archaeological<br />

Interpretation:<strong>The</strong> Pottery <strong>of</strong> the Maghreb. In F. R. Matson (ed) Ceramics and<br />

Man. New York: Viking,<br />

Bath, M., Deguara J., 2003. Southern Hemisphere Tropical Cyclones: Historical Tracks<br />

and Data. Electronic maps prepared using Australian Bureau <strong>of</strong> Meteorology<br />

data. http://australiasevereweather.com/cyclones/history.htm<br />

Baxter, M. J., 1994. Exploratory Multivariate Analysis in Archaeology. Edinburgh<br />

510


<strong>University</strong> Press.<br />

Bedford, S., 1999. Lapita and Post-Lapita Ceramic Sequences from Erromango, Southern<br />

Vanuatu. In J. C. Galipaud and I. Lilley (eds) <strong>The</strong> Pacific, 5000-2000 BP:<br />

Colonizations and Transformations. Paris: IRD Editions, pp.127-37.<br />

---, 2000. Pieces <strong>of</strong> the Vanuatu Puzzle: Archaeology <strong>of</strong> the North, South and Centre.<br />

PhD thesis. Department <strong>of</strong> Archaeology and Natural History, Research School <strong>of</strong><br />

Pacific and Asian Studies, Australian National <strong>University</strong>.<br />

---, 2002. Personal Communication: to M. Felgate. <strong>Auckland</strong><br />

Bellwood, P., 1978. Man's Conquest <strong>of</strong> the Pacific. Sydney: Collins.<br />

Best, S., 1984. Lakeba: <strong>The</strong> Prehistory <strong>of</strong> a Fijian Island. PhD thesis. Anthropology<br />

Department, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

---, 2002. Lapita: A View from the East. New Zealand Archaeological Association<br />

Monograph 24. <strong>Auckland</strong>.<br />

Bickler, S. H., 1998. Eating Stone and Dying: Archaeological Survey on Woodlark<br />

Island, Milne Bay Province, Papua New Guinea. PhD thesis. Department <strong>of</strong><br />

Anthropology, <strong>University</strong> <strong>of</strong> Virginia.<br />

Binford, L. R., 1962. Archaeology as Anthropology. American Antiquity 44: 455-70<br />

---, 1964. A Consideration <strong>of</strong> Archaeological Research Design. American Antiquity 29:<br />

425-41<br />

Blankholm, H. P., 1991. Intrasite Spatial Analysis in <strong>The</strong>ory and Practice. Aarhus:<br />

Aarhus <strong>University</strong> Press.<br />

Braun, D. P., 1983. Pots as Tools. In J. A. Moore and A. S. Keene (eds) Archaeological<br />

Hammers and <strong>The</strong>ories. New York, London: Academic Press, pp.107-36.<br />

Brenchley, P. J., Harper D. T., 1998. Paleoecology: Ecosystems, Environments and<br />

Evolution. Chapman and Hall.<br />

Bruns, T. R., Cooper A. K., Mann D., Vedder J. G., 1986. Seismic Stratigraphy and<br />

Structure <strong>of</strong> Sedimentary Basins in the Solomon Islands Region. In J. G. Vedder<br />

511


(ed) Geology and Offshore Resources <strong>of</strong> Pacific Island Arcs - Central and<br />

Western Solomon Islands. AAPG Circum-Pacific Council for Energy and Mineral<br />

Resources, 4, pp.177-214.<br />

Bunge, J., Fitzpatrick M., 1993. Estimating the Number <strong>of</strong> Species: A Review. Journal <strong>of</strong><br />

the American Statistical Association 88 (421): 364-73<br />

Burley, D. V., 1994. Settlement Pattern and Tongan Prehistory. Journal <strong>of</strong> the Polynesian<br />

Society 103 (4): 379-411<br />

Burley, D. V., Dickinson W. R., 2001. Origin and Significance <strong>of</strong> a Founding Settlement<br />

in Polynesia. Proceedings <strong>of</strong> the National Academy <strong>of</strong> Sciences 98 (20): 11829-31<br />

Chao, A., 1984. Nonparametric Estimation <strong>of</strong> the Number <strong>of</strong> Classes in a Population.<br />

Scandinavian Journal <strong>of</strong> Statistics 11: 265-70<br />

Chao, A., Lee S.-M., 1992. Estimating the Number <strong>of</strong> Classes Via Sample Coverage.<br />

Journal <strong>of</strong> the American Statistical Association 87 (417): 210-7<br />

Clark, G. R., 1999. Post-Lapita Fiji: Cultural Transformation in the Mid-Sequence. PhD<br />

thesis. Department <strong>of</strong> Archaeology and Natural History, Research School <strong>of</strong><br />

Pacific and Asian Studies, Australian National <strong>University</strong>.<br />

Clarke, D. L., 1968. Analytical Archaeology. London: Methuen.<br />

Colwell, R. K., 1997. Estimates: Statistical Estimation <strong>of</strong> Species Richness and Shared<br />

Species from Samples. Connecticut: http://viceroy.eeb.uconn.edu/estimates.<br />

Cowgill, G. L., 1972. Models, Methods and Techniques for Seriation. In D. L. Clarke (ed)<br />

Models in Archaeology. London: Methuen, pp.381-424.<br />

---, 1975. A Selection <strong>of</strong> Samplers: Comments on Archaeostatistics. In J. W. Mueller (ed)<br />

Sampling in Archaeology. Tucson: <strong>The</strong> <strong>University</strong> <strong>of</strong> Arizons Press, pp.258-74.<br />

Cowgill, G. L., Altshul J. H., Sload R. S., 1984. Spatial Analysis <strong>of</strong> Teotihuacan: A<br />

Mesoamerican Metropolis. In H. Hietala (ed) Intrasite Spatial Analysis in<br />

Archaeology. Cambridge <strong>University</strong> Press, pp.154-95.<br />

Davenport, W., 1972. Preliminary Excavations on Santa Ana Island, Eastern Solomon<br />

512


Islands. Archaeology and Physical Anthropology in Oceania 7 (3): 165-83<br />

David, B., Wilson M., 1999. Re-Reading the Landscape: Place and Identity in NE<br />

Australia During the Late Holocene. Cambridge Archaeological Journal 9 (2):<br />

163-88<br />

David, N., 1972. On the Life Span <strong>of</strong> Pottery, Type Frequencies, and Archaeological<br />

Inference. American Antiquity 37 (1): 141-2<br />

Davis, W. M., 1928. <strong>The</strong> Coral Reef Problem. American Geographic Society Special<br />

Publication 9.<br />

Day, T. J., 1980. A Study <strong>of</strong> the Transport <strong>of</strong> Graded Sediments. Wallingford, England :<br />

Hydraulics Research Station.<br />

de Barros, P., 1982. <strong>The</strong> Effects <strong>of</strong> Variable Site Occupation Span on the Results <strong>of</strong><br />

Frequency Seriation. American Antiquity 47 (2): 291-315<br />

De Boer, W. R., 1983. <strong>The</strong> Archaeological Assemblage as Preserved Death Assemblage.<br />

In J. A. Moore and A. S. Keene (eds) Archaeological Hammers and <strong>The</strong>ories.<br />

New York: Academic Press,<br />

Dickinson, W. R., 2000a. Petrography <strong>of</strong> Sand Tempers in Prehistoric Sherds from<br />

Roviana Lagoon, New Georgia, Solomon Islands. Tucson: Unpublished Report,<br />

<strong>University</strong> <strong>of</strong> Arizona, Geology Department, pp.10.<br />

---, 2000b. Petrography <strong>of</strong> Sand Tempers in Prehistoric Watom Sherds and Comparison<br />

with Other Temper Suites <strong>of</strong> the Bismarck Archipelago. New Zealand Journal <strong>of</strong><br />

Archaeology 20 (1998): 161-82<br />

Dickinson, W. R., Green R. C., 1998. Geoarchaeological Context <strong>of</strong> Holocene Subsidence<br />

at the Ferry Berth Lapita Site, Mulifanua, Upolu, Samoa. Geoarchaeology 13 (3):<br />

239-67<br />

Donovan, L. J., 1973. A Study <strong>of</strong> the Decorative System <strong>of</strong> the Lapita Potters in Reefs and<br />

Santa Cruz Islands. MA thesis. Department <strong>of</strong> Anthropology, <strong>University</strong> <strong>of</strong><br />

<strong>Auckland</strong>.<br />

513


Duff, A. I., 1996. Ceramic Micro-Seriation: Types or Attributes? American Antiquity 61<br />

(1): 89-101<br />

Dunkley, P. N., 1986. <strong>The</strong> Geology <strong>of</strong> the New Georgia Group, Western Solomon Islands<br />

British Geological Survey Overseas Directorate, Keyworth, Nottinghamshire<br />

Dunnell, R. C., 1970. Seriation Method and Its Evaluation. American Antiquity 35 (3):<br />

305-19<br />

---, 1971. Sabl<strong>of</strong>f and Smith's "the Importance <strong>of</strong> Both Analytic and Taxonomic<br />

Classification in the Type-Variety System". American Antiquity 36 (1): 115-8<br />

---, 1978. Style and Function: A Fundamental Dichotomy. American Antiquity 43: 192-<br />

202<br />

---, 1986. Methodological Issues in Americanist Artifact Classification. In M. B. Schiffer<br />

(ed) Advances in Archaeological Method and <strong>The</strong>ory. Academic Press, pp.149-<br />

207.<br />

---, 2001. Foreword. In T. D. Hurt and G. F. M. Rakita (eds) Style and Function:<br />

Conceptual Issues in Evolutionary Archaeology. Westport, Connecticutt: Bergin<br />

and Harvey, pp.xiii-xxiv.<br />

Egl<strong>of</strong>f, B. J., 1971. Collingwood Bay and the Trobriand Islands in Recent Prehistory.<br />

PhD thesis. Australian National <strong>University</strong>.<br />

---, 1973. A Method for Counting Ceramic Rim Sherds. American Antiquity 38 (3): 351-3<br />

---, 1979. Recent Prehistory in Southeast Papua. Terra Australis. 4, Canberra:<br />

Department <strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific Studies, ANU.<br />

Ericson, J. E., Read D. W., Burke C., 1972. Research Design: <strong>The</strong> Relationships between<br />

the Primary Functions and the Physical Properties <strong>of</strong> Ceramic Vessels and <strong>The</strong>ir<br />

Implications for Distributions on an Archaeological Site. Anthropology UCLA 3<br />

(2): 84-95<br />

Feathers, J. K., 1997. <strong>The</strong> Application <strong>of</strong> Luminescence Dating in American Archaeology.<br />

Journal <strong>of</strong> Archaeological Method and <strong>The</strong>ory 4 (1): 1-67<br />

514


---, 2002. Luminescence Dating <strong>of</strong> Pottery Sherds from the Solomon Islands. Unpublished<br />

Analysis Report, Seattle: <strong>University</strong> <strong>of</strong> Washington.<br />

Felgate, M. W., 2002. A Roviana Ceramic Sequence and the Prehistory <strong>of</strong> Near Oceania:<br />

Work in Progress. In G. R. Clark, A. J. Anderson and T. Vunidilo (eds) <strong>The</strong><br />

Archaeology <strong>of</strong> the Lapita Dispersal in Oceania: Papers from the 4th Lapita<br />

Conference, June 2000, Canberra, Australia. Pandanus Press, Research School<br />

<strong>of</strong> Pacific and Asian studies, ANU, pp.39-60.<br />

Felgate, M. W., Bickler S. H., n.d. Pots as Species: Estimating a Parent Population <strong>of</strong><br />

Vessels from a Sherd Sample. Unpublished Manuscript in the Posession <strong>of</strong> the<br />

Authors.<br />

Felgate, M. W., Dickinson W. R., 2001. Late-Lapita and Post-Lapita Pottery Transfers:<br />

Evidence from Intertidal-Zone Sites <strong>of</strong> Roviana Lagoon, Western Province,<br />

Solomon Islands. In M. Jones and P. J. Sheppard (eds) Proceedings <strong>of</strong> the 2001<br />

Austalasian Archaeolmentry Conference. <strong>Auckland</strong>: Research Papers in<br />

Anthropology and Linguistics, pp.105-22.<br />

Ferrari, B., Adams J., 1990. Biogenetic Modifications <strong>of</strong> Marine Sediments and <strong>The</strong>ir<br />

Influence on Archaeological Material. International Journal <strong>of</strong> Nautical<br />

Archaeology and Underwater Exploration 19: 139-51<br />

Flemming, N. C., 1983. Survival <strong>of</strong> Submerged Lithic and Bronze Age Artifact Sites: A<br />

Review. In P. M. Masters and N. C. Flemming (eds) Quaternary Coastlines and<br />

Marine Archaeology. Academic Press, pp.135-73.<br />

Fontana, V., 1998. Procedures to Analyse Intra-Site Pottery Distribution, Applied to the<br />

Neolithic Site <strong>of</strong> Fimon, Molino Casarotto (Italy), Housesite No.3. Journal <strong>of</strong><br />

Archaeological Science 25 (11): 1067-72<br />

Frost, E. L., 1974. Archaeological Excavations <strong>of</strong> Fortified Sites on Tavenui, Fiji. Asian<br />

and Pacific Archaeology Series. 6, Honolulu: Social Science Research Institute,<br />

<strong>University</strong> <strong>of</strong> Hawaii.<br />

515


Gaffney, V. L., Bintcliff J., Slapsak B., 1991. Site Formation Processes and the Hvar<br />

Survey Project, Yugoslavia. In A. J. Sch<strong>of</strong>ield (ed) Interpreting Atrefact Scatters:<br />

Contributions to Ploughzone Archaeology. Oxford: Oxbow Books,<br />

Galipaud, J.-C., 1996. New Caledonia: Some Recent Archaeological Perspectives. In J.<br />

M. Davidson, G. Irwin, B. F. Leach, A. Pawley and D. Brown (eds) Oceanic<br />

Culture History: Essays in Honour <strong>of</strong> Roger Green. New Zealand Journal <strong>of</strong><br />

Archaeology Special Publication, pp.297-305.<br />

---, 1998. <strong>The</strong> Lapita Site <strong>of</strong> Atanoasao: Malo, Vanuatu ORSTOM, Port Vila<br />

Gardin, J. C., 1967. Methods for the Descriptive Analysis <strong>of</strong> Archaeological Material.<br />

American Antiquity 32: 13-30<br />

Gell, A., 1992. <strong>The</strong> Technology <strong>of</strong> Enchantment and the Enchantment <strong>of</strong> Technology. In<br />

J. Coote and A. Shelton (eds) Anthropology, Art and Aesthetes. Oxford:<br />

Clarendon Press, pp.40-63.<br />

Gifford, E. W., Shutler R. J., 1956. Archaeological Excavations in New Caledonia.<br />

Anthropological Records. 18(1), Berkeley and Los Angeles: <strong>University</strong> <strong>of</strong><br />

California Press.<br />

Goldmann, K., 1971. Some Archaeological Criteria for Chronological Seriation. In F. R.<br />

Hodson, D. G. Kendall and P. Tautu (eds) Mathematics in the Archaeological and<br />

Information Sciences. Edinburgh: Edinburgh <strong>University</strong> Press,<br />

Golson, J., 1971. Lapita Ware and Its Transformations. In R. C. Green and M. Kelly (eds)<br />

Studies in Oceanic Culture History. Honolulu: Department <strong>of</strong> Anthropology,<br />

Bernice P. Bishop Museum, pp.67-76.<br />

---, 1991. Two Sites at Lasigi, New Ireland. In J. Allen and C. Gosden (eds) Report <strong>of</strong> the<br />

Lapita Homeland Project. Canberra, Australia: Department <strong>of</strong> Prehistory,<br />

Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian National <strong>University</strong>, pp.244-59.<br />

---, 1992. <strong>The</strong> Pottery from Lasigi, New Ireland. Presented at Poterie Lapita et<br />

516


Peuplement: Actes Du Colloque Lapita, Janvier 1992, Noumea, New Caledonia<br />

Gorecki, P., 1985. Test Excavation at Lemau, New Ireland Province and Report <strong>of</strong> an<br />

Excursion to Langovai, New Hanover, New Ireland Province. In J. Allen (ed)<br />

Lapita Homeland Project: Report <strong>of</strong> the 1985 Field Season. Melbourne: La Trobe<br />

<strong>University</strong>, pp.20-2.<br />

---, 1992. A Lapita Smoke Screen ? In J. C. Galipaud (ed) Poterie Lapita Et Peuplement:<br />

Actes Du Colloque Lapita, Janvier 1992. Noumea, New Caledonia:<br />

O.R.S.T.R.O.M., pp.27-47.<br />

Gosden, C., 1985. An Archaeological Survey <strong>of</strong> the Arawe Islands, West New Britain.<br />

Lapita Homeland Project: Report <strong>of</strong> the 1985 Field Season. Melbourne: La Trobe<br />

<strong>University</strong>, pp.54-5.<br />

---, 1989. Prehistoric Social Landscapes <strong>of</strong> the Arawe Islands, West New Britain Province,<br />

Papua New Guinea. Archaeology in Oceania 24 (2): 45-58<br />

---, 1991a. Learning About Lapita in the Bismarck Archipelago. In J. Allen and C. Gosden<br />

(eds) Report <strong>of</strong> the Lapita Homeland Project. Canberra, Australia: Department <strong>of</strong><br />

Prehistory, Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian National <strong>University</strong>,<br />

pp.260-8.<br />

---, 1991b. Towards an Understanding <strong>of</strong> the Regional Archaeological Record from the<br />

Arawe Islands, West New Britain, Papua New Guinea. In J. Allen and C. Gosden<br />

(eds) Report <strong>of</strong> the Lapita Homeland Project. Canberra, Australia: Department <strong>of</strong><br />

Prehistory, Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian National <strong>University</strong>,<br />

pp.205-16.<br />

Gosden, C., Allen J., Ambrose W. R., Anson D., Golson J., et al, 1989. Lapita Sites <strong>of</strong> the<br />

Bismarck Archipelago. Antiquity 63: 561-86<br />

Gosden, C., Head L., 1994. Landscape - a Usefully Ambiguous Concept. Archaeology in<br />

Oceania 29 (3): 113-6<br />

Gosden, C., Webb J., 1994. <strong>The</strong> Creation <strong>of</strong> a Papua-New Guinean Landscape:<br />

517


Archaeological and Geomorphological Evidence. Journal <strong>of</strong> Field Archaeology<br />

21: 29-51<br />

Gould, S. J., Vbra E., 1982. Exaptation- a Missing Term in the Science <strong>of</strong> Form.<br />

Paleobiology 8: 4-15<br />

Graves, M. W., 1998. <strong>The</strong> History <strong>of</strong> Method and <strong>The</strong>ory in the Study <strong>of</strong> Prehistoric<br />

Puebloan Pottery Style in the American Southwest. Journal <strong>of</strong> Archaeological<br />

Method and <strong>The</strong>ory 5 (4): 309-43<br />

Graves, M. W., Cachola-Abad C. K., 1996. Seriation as a Method <strong>of</strong> Chronologically<br />

Ordering Architectural Design Traits: An Example from Hawai'i. Archaeology in<br />

Oceania 31 (1): 19-32<br />

Green, R. C., 1974a. Excavation <strong>of</strong> the Prehistoric Occupations <strong>of</strong> Su-Sa-3. In R. C.<br />

Green and J. Davidson (eds) Archaeology in Western Samoa: Volume II.<br />

<strong>Auckland</strong>: <strong>Auckland</strong> Institute and Museum, pp.108-62.<br />

---, 1974b. Review <strong>of</strong> Portable Artefacts from Western Samoa. In R. C. Green and J.<br />

Davidson (eds) Archaeology in Western Samoa: Volume II. pp.245-74.<br />

---, 1976a. Lapita Sites in the Santa Cruz Group. In R. C. Green and M. M. Cresswell<br />

(eds) Southeast Solomons Culture History. Wellington, New Zealand: <strong>The</strong> Royal<br />

Society <strong>of</strong> New Zealand, Bulletin 11, pp.245-65.<br />

---, 1976b. A Late Prehistoric Sequence from Su' Ena Village, Uki. In R. C. Green and M.<br />

M. Cresswell (eds) Southeast Solomons Culture History. Wellington, New<br />

Zealand: <strong>The</strong> Royal Society <strong>of</strong> New Zealand, Bulletin 11, pp.181-91.<br />

---, 1976c. A Late-Prehistoric Settlement in Star Harbour. In R. C. Green and M. M.<br />

Cresswell (eds) Southeast Solomons Culture History. Wellington, New Zealand:<br />

<strong>The</strong> Royal Society <strong>of</strong> New Zealand, Bulletin 11, pp.133-47.<br />

---, 1978. New Sites with Lapita Pottery and <strong>The</strong>ir Implications for an Understanding <strong>of</strong><br />

the Settlement <strong>of</strong> the Western Pacific. Working Papers in Anthropology,<br />

Archaeology, Linguistics and Maori Studies. 51, <strong>Auckland</strong>, New Zealand:<br />

518


Department <strong>of</strong> Anthropology, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

---, 1979. Lapita. In J. D. Jennings (ed) <strong>The</strong> Prehistory <strong>of</strong> Polynesia. Cambridge,<br />

Massachusetts: Harvard <strong>University</strong> Press, pp.27-59.<br />

---, 1990. Lapita Design Analysis: <strong>The</strong> Mead System and Its Use; a Potted History.<br />

Presented at Lapita Design, Form and Composition: Proceedings <strong>of</strong> the Lapita<br />

Design Workshop, Canberra, Australia, December 1988, Canberra Australia<br />

---, 1991a. <strong>The</strong> Lapita Cultural Complex: Current Evidence and Proposed Models.<br />

Bulletin <strong>of</strong> the Indo-Pacific Prehistory Association 11: 295-305<br />

---, 1991b. Near and Remote Oceania -- Disestablishing "Melanesia" in Culture History.<br />

In A. Pawley (ed) Man and a Half: Essays in Pacific Anthropology and<br />

Ethnobiology in Honour <strong>of</strong> Ralph Bulmer. <strong>Auckland</strong>: <strong>The</strong> Polynesian Society,<br />

pp.491-502.<br />

---, 1991c. A Reappraisal <strong>of</strong> the Dating for Some Lapita Sites in the Reef/Santa Cruz<br />

Group <strong>of</strong> the Southeast Solomons. Journal <strong>of</strong> the Polynesian Society 100: 197-<br />

207<br />

---, 1992. Definitions <strong>of</strong> the Lapita Cultural Complex and Its Non-Ceramic Component.<br />

in Poterie Lapita et Peuplement: Actes Du Colloque Lapita, Janvier 1992, pp. 7-<br />

21. Noumea, New Caledonia: O.R.S.T.R.O.M.<br />

---, 1994. Prehistoric Transfers <strong>of</strong> Portable Items During the Lapita Horizon in Remote<br />

Oceania: A Review. Ippa 15th Congress: Chiang Mai -Jan. 5-12.<br />

Green, R. C., Anson D., 1991. <strong>The</strong> Reber-Rakival Site on Watom. Implications <strong>of</strong> the<br />

1985 Excavations at the SAC and SDI Localities. In J. Allen and C. Gosden (eds)<br />

Report <strong>of</strong> the Lapita Homeland Project. Canberra, Australia: Department <strong>of</strong><br />

Prehistory, Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian National <strong>University</strong>,<br />

pp.170-81.<br />

---, 2000a. Archaeological Investigations on Watom Island: Early Work, Outcomes <strong>of</strong><br />

Recent Investigations, and Future Prospects. New Zealand Journal <strong>of</strong> Archaeology<br />

519


20 (1998): 183-97<br />

---, 2000b. Excavations at Kainapirina (SAC), Watom Island, Papua New Guinea. New<br />

Zealand Journal <strong>of</strong> Archaeology 20 (1998): 29-94<br />

Green, R. C., Davidson J. M., 1969. Description and Classification <strong>of</strong> Samoan Adzes. In<br />

R. C. Green and J. M. Davidson (eds) Archaeology in Western Samoa. <strong>Auckland</strong>:<br />

<strong>Auckland</strong> Institute and Museum, pp.21-32.<br />

Hagstrum, M. B., Hildebrand J. A., 1990. <strong>The</strong> Two-Curvature Method for Reconstructing<br />

Ceramic Morphology. American Antiquity 55 (2): 388-403<br />

Hamilton, D., Woodward R., 1984. A Sunken 17th Century City: Port Royal, Jamaica.<br />

Archaeology 37: 38-45<br />

Hedrick, J. D., n.d. Archaeological Investigation <strong>of</strong> Malo Prehistory:Lapita Settlement<br />

Strategy in the Northern New Hebrides. Unfinished draft PhD thesis. <strong>University</strong><br />

<strong>of</strong> Pennsylvania.<br />

Henderson, A. G., 1998. Middle Fort Ancient Villages and Organizational Complexity<br />

in Central Kentucky. PhD thesis. university <strong>of</strong> kentucky.<br />

Henderson, H. M., 1992. <strong>The</strong> Strickler Site: <strong>The</strong> Analysis <strong>of</strong> a Nineteenth to Twentieth<br />

Century Urban Farmstead in Markham County, Ontario (Canada). MA thesis.<br />

<strong>University</strong> <strong>of</strong> Toronto.<br />

Hendren, G. H., 1976. Recent Settlement Pattern Changes on Ulawa, Southeast Solomon<br />

Islands. In R. C. Green and M. M. Cresswell (eds) Southeast Solomons Culture<br />

History. Wellington, New Zealand: <strong>The</strong> Royal Society <strong>of</strong> New Zealand, Bulletin<br />

11, pp.149-59.<br />

Henrickson, E. F., 1990. Investigating Ancient Ceramic Form and Use. In W. D. Kingery<br />

(ed) <strong>The</strong> Changing Roles <strong>of</strong> Ceramics in Society: 26,000 B.P. To the Present.<br />

Westerville, Ohio: <strong>The</strong> American Ceramic Society, Inc., pp.83-117.<br />

Henrickson, E. F., Mcdonald M. A., 1985. Ceramic Form and Function: An Ethnographic<br />

Search and Archaeological Application. American Anthropologist 85: 630-43<br />

520


Hodder, I., Orton C., 1976. Spatial Analysis in Archaeology. New Studies in<br />

Archaeology. Cambridge <strong>University</strong> Press.<br />

Hydrographic-Office, G. B., 1971. Pacific Islands Pilot Volume I: <strong>The</strong> Western Groups<br />

Comprising the Territory <strong>of</strong> Papua Including the Louisiade Archipelago, the<br />

North-East and North Coasts <strong>of</strong> New Guinea, the Solomon Islands, the Bismarck<br />

Archipelago, and the Caroline and Marianas Islands. Hydrographer <strong>of</strong> the Navy.<br />

Irwin, G., 1972. An Archaeological Survey in the Shortland Islands, B.S.I.P. MA thesis.<br />

Anthropology Department, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

---, 1985. <strong>The</strong> Emergence <strong>of</strong> Mailu: As a Central Place in Coastal Papuan Prehistory.<br />

Terra Australis. 10, Canberra: Department <strong>of</strong> Prehistory, Research School <strong>of</strong><br />

Pacific Studies, ANU.<br />

---, 1992. <strong>The</strong> Prehistoric Exploration and Colonization <strong>of</strong> the Pacific. Cambridge<br />

<strong>University</strong> Press.<br />

Ishimura, T., 2002. In the Wake <strong>of</strong> Lapita: Transformation <strong>of</strong> Lapita Designs and Gradual<br />

Dispersal <strong>of</strong> the Lapita Peoples. People and Culture in Oceania 18: 77-97<br />

Jensen, K., Jensen J., Clegg C., 1999. Inferring Intensity <strong>of</strong> Site Use from the Breakdown<br />

Rate and Discard Patterns <strong>of</strong> Fire-Cracked Rock at Playa View Dune. Utah<br />

Archaeology 12: 51-64<br />

Kintigh, K. W., 1984. Measuring Archaeological Diversity by Comparison with Simulated<br />

Assemblages. American Antiquity 49 (1): 44-54<br />

---, 1990. Intrasite Spatial Analysis: A Commentary on Major Methods. In A. Voorrips<br />

(ed) Mathematics and Information Science in Archaeology: A Flexible<br />

Framework. Bonn: Holos, pp.165-200.<br />

Kirch, P. V., 1985. Archaeological Fieldwork on Eloaua and Emananus Islands, St.<br />

Matthias Group, New Ireland Province. In J. Allen (ed) Lapita Homeland Project:<br />

Report <strong>of</strong> the 1985 Field Season. Melbourne: La Trobe <strong>University</strong>, pp.14-9.<br />

---, 1987a. Is <strong>The</strong>re an Early Far Western Lapita Province? Sample Size Effects and New<br />

521


Evidence from Eloaua Island. Archaeology in Oceania 22: 123-6<br />

---, 1987b. Lapita and Oceanic Cultural Origins: Excavations in the Mussau Islands,<br />

Bismarck Archipelago, 1985. Journal <strong>of</strong> Field Archaeology 14: 163-80<br />

---, 1988a. Niuatopupapu: <strong>The</strong> Prehistory <strong>of</strong> a Polynesian Chiefdom. Monograph 5,<br />

Seattle: Burke Museum, Seattle.<br />

---, 1988b. Problems and Issues in Lapita Archaeology. In P. V. Kirch and T. L. Hunt<br />

(eds) Archaeology <strong>of</strong> the Lapita Cultural Complex: A Critical Review. Seattle:<br />

Thomas Burke Memorial Washington State Museum Research Report No. 5,<br />

pp.167-5.<br />

---, 1988c. <strong>The</strong> Talepakemalai Lapita Site and Oceanic Prehistory. National Geographic<br />

Research 4 (3): 38-342<br />

---, 1997. <strong>The</strong> Lapita Peoples: Ancestors <strong>of</strong> the Oceanic World. Blackwell.<br />

---, 2001. Lapita and Its Transformations in near Oceania: Archaeological Investigations<br />

in the Mussau Islands, Papua New Guinea, 1985-88. Berkeley: <strong>University</strong> <strong>of</strong><br />

California.<br />

Kirch, P. V., Green R. C., 2001. Hawaiki, Ancestral Polynesia: An Essay in Historical<br />

Anthropology. Cambridge <strong>University</strong> Press.<br />

Kirch, P. V., Hunt T. L., 1988. <strong>The</strong> Spatial and Temporal Boundaries <strong>of</strong> Lapita. In P. V.<br />

Kirch and T. L. Hunt (eds) Archaeology <strong>of</strong> the Lapita Cultural Complex: A<br />

Critical Review. Seattle: Thomas Burke Memorial Washington State Museum<br />

Research Report No. 5, pp.9-31.<br />

Kirch, P. V., Hunt T. L., Weisler M., Butler V., Allen M. S., 1991. Mussau Islands<br />

Prehistory: Results <strong>of</strong> the 1985-1986 Excavations. In J. Allen and C. Gosden (eds)<br />

Report <strong>of</strong> the Lapita Homeland Project. Canberra, Australia: Department <strong>of</strong><br />

Prehistory, Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian National <strong>University</strong>,<br />

pp.144-63.<br />

Kirch, P. V., Rosendahl P., 1973. Archaeological Investigation <strong>of</strong> Anuta. In D. E. Yen and<br />

522


J. Gordon (eds) Anuta: A Polynesian Outlier in the Solomon Islands. Honolulu:<br />

Bishop Museum,<br />

Krantz, G. S., 1968. A New Method for Counting Mammal Bones. American Journal <strong>of</strong><br />

Archaeology 72: 286-8<br />

Lampert, R. J., 1966. Archaeological Reconnaissance in Papua and New Guinea.<br />

Canberra: Unpublished Mimeograph, Department <strong>of</strong> Prehistory, Research School<br />

<strong>of</strong> Pacific Studies, ANU.<br />

Le Blanc, S. A., 1975. Micro-Seriation: A Method for Chronologic Differentiation.<br />

American Antiquity 40 (1): 22-38<br />

Lep<strong>of</strong>sky, D., 1988. <strong>The</strong> Environmental Context <strong>of</strong> Lapita Settlement Locations. In P. V.<br />

Kirch and T. L. Hunt (eds) Archaeology <strong>of</strong> the Lapita Cultural Complex: A<br />

Critical Review. Seattle: Burke Museum, Seattle, pp.33-48.<br />

Lilley, I., 1985. Archaeological Fieldwork in Duke <strong>of</strong> York Islands, East New Britain<br />

Province. In J. Allen (ed) Lapita Homeland Project: Report <strong>of</strong> the 1985 Field<br />

Season. Melbourne: La Trobe <strong>University</strong>, pp.41-2.<br />

---, 1986. Prehistoric Exchange in the Vitiaz Strait, Papua New Guinea. PhD thesis.<br />

Australian National <strong>University</strong>.<br />

---, 1991. Lapita Sites in the Duke <strong>of</strong> York Islands. In J. Allen and C. Gosden (eds) Report<br />

<strong>of</strong> the Lapita Homeland Project. Canberra, Australia: Department <strong>of</strong> Prehistory,<br />

Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian National <strong>University</strong>, pp.164-9.<br />

Linton, R. E., 1925. Archaeology <strong>of</strong> the Marquesas Islands. Bernice P Bishop Museum<br />

Bulletin 23. Honolulu: Bayard Dominick Expedition Bulletin Number 10.<br />

---, 1944. North American Cooking Pots. American Antiquity IX (4): 369-80<br />

Longacre, W. A., 1991. Sources <strong>of</strong> Ceramic Variability among the Kalinga <strong>of</strong> Northern<br />

Luzon. In W. A. Longacre (ed) Ceramic Ethnoarchaeology. Tucson: <strong>University</strong><br />

<strong>of</strong> Arizona Press, pp.95-111.<br />

Lyman, R. L., Wolverton S., O'Brien M. J., 1998. Seriation, Superposition, and<br />

523


Interdigitation: A History <strong>of</strong> Americanist Graphic Depictions <strong>of</strong> Culture Change.<br />

American Antiquity 63 (2): 239-61<br />

Macmillan, D. H., 1966. Tides. New York: American Elsevier Publishing Company Inc.<br />

Mann, P., Taylor F. W., Lagoe M. B., Quarles A., Burr G., 1998. Accelerating Late<br />

Quaternary Uplift <strong>of</strong> the New Georgia Island Group (Solomon Island Arc) in<br />

Response to Subduction <strong>of</strong> the Recently Active Woodlark Spreading Centre and<br />

Coleman Seamount. Tectonophysics 295: 259-306<br />

Marquardt, W. H., 1978. Advances in Archaeological Seriation. Advances in<br />

Archaeological Method and <strong>The</strong>ory. Journal <strong>of</strong> Archaeological Method and<br />

<strong>The</strong>ory, pp.257-314.<br />

Mccoy, P. C., Cleghorn P. L., 1988. Archaeological Investigations on Santa Cruz<br />

(Nendo), Southeast Solomon Islands: Summary Report. Archaeology in Oceania<br />

23: 104-15<br />

McNutt, C. H., 1973. On the Methodological Validity <strong>of</strong> Frequency Seriation. American<br />

Antiquity 38 (1): 45-60<br />

Mead, S. M., 1975. <strong>The</strong> Decorative System <strong>of</strong> the Lapita Potters <strong>of</strong> Sigatoka, Fiji. In S.<br />

M. Mead, H. Birks, L. Birks and E. Shaw (eds) <strong>The</strong> Lapita Pottey Style <strong>of</strong> Fiji and<br />

Its Associations. Wellington: <strong>The</strong> Polynesian Society, pp.19-43.<br />

Mead, S. M., Birks L., Birks H., Shaw E., 1975. <strong>The</strong> Lapita Pottery Style <strong>of</strong> Fiji and Its<br />

Associations. Wellington: <strong>The</strong> Polynesian Society.<br />

Miller, D., 1985. Artefacts as Categories: A Study <strong>of</strong> Ceramic Variability in Central<br />

India. New Studies in Archaeology. London: Cambridge <strong>University</strong> Press.<br />

Miller, D., Roe D., 1982. <strong>The</strong> Solomon Islands National Sites Survey: <strong>The</strong> First Phase.<br />

Bulletin <strong>of</strong> the Indo-Pacific Prehistory Association 3: 47-51<br />

Mills, B. J., 1989. Integrating Functional Analyses <strong>of</strong> Vessels and Sherds through Models<br />

<strong>of</strong> Ceramic Assemblage Formation. World Archaeology 21 (1): 133-47<br />

Nojima, Y., 2002. Personal Communication: to M. Felgate.<br />

524


O'Brien, M. J., Lyman R. L., 2000. Seriation, Stratigraphy and Index Fossils: <strong>The</strong><br />

Backbone <strong>of</strong> Archaeological Dating. New York: Kluwer Academic/Plenum.<br />

Open <strong>University</strong> Oceanography Team, 1989. Waves, Tides and Shallow-Water Processes.<br />

Oxford: Pergamon Press.<br />

Oreihaka, E., 1997. Freshwater and Marine Aquatic Resources in Solomon Islands.<br />

Orton, C., 1982. Computer Simulation Experiments to Assess the Performance <strong>of</strong><br />

Measures <strong>of</strong> Quantity <strong>of</strong> Pottery. World Archaeology 14 (1): 1-20<br />

---, 1993. How Many Pots Make Five?- an Historical Review <strong>of</strong> Pottery Quantification.<br />

Archaeometry 35 (2): 169-84<br />

---, 2000. Sampling in Archaeology. Cambridge Manuals in Archaeology. Cambridge<br />

<strong>University</strong> Press.<br />

Orton, C., Tyers P., Vince A., 1993. Pottery in Archaeology. Cambridge Manuals in<br />

Archaeology. Cambridge <strong>University</strong> Press.<br />

Orton, C., Tyers P. A., 1991. Counting Broken Objects: <strong>The</strong> Statistics <strong>of</strong> Ceramic<br />

Assemblages. Proceedings <strong>of</strong> the British Academy 77: 163-84<br />

Parker, V. N. M., 1981. Vessel Forms <strong>of</strong> the Reef Islands Se-Rf-2 Site and <strong>The</strong>ir<br />

Relationships to Vessel Forms in Other Western Lapita Sites <strong>of</strong> the Reef/Santa<br />

Cruz and Island Melanesian Area. MA thesis. Anthropology Department,<br />

<strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

Petchey, F. J., 1995. <strong>The</strong> Archaeology <strong>of</strong> Kudon: Archaeological Analysis <strong>of</strong> Lapita<br />

Ceramics from Mulifanua, Samoa, and Sigatoka, Fiji. MA thesis. Anthropology<br />

Department, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

Phelan, M., 1997. Scratching the Surface: <strong>The</strong> Lapita Pottery <strong>of</strong> Makegur, Papua New<br />

Guinea. MA thesis. School <strong>of</strong> archaeology, La Trobe <strong>University</strong>.<br />

Plog, S., 1985. Estimating Vessel Orifice Diameters: Methods and Measurement Error. In<br />

B. A. Nelson (ed) Decoding Prehistoric Ceramics. Carbondale and Edwardsville:<br />

Southern Illinois <strong>University</strong> Press, pp.243-53.<br />

525


Poulsen, J. I., Polach H. A., 1972. Outlier Archaeology: Bellona. A Preliminary Report on<br />

Field Work and Radiocarbon Dates. Archaeology and Physical Anthropology in<br />

Oceania 7: 184-214<br />

Prior, C., 2000. Personal Communication: Rafter Radiocarbon Laboratory, Institute <strong>of</strong><br />

Geological and Nuclear Sciences, Lower Hutt, New Zealand, to M. Felgate.<br />

Rathje, R., 2000. Lapita Development as the First Holiday Inns. In A. Anderson and T.<br />

Murray (eds) Australian Archaeologist: Collected Papers in Honour <strong>of</strong> Jim Allen.<br />

Canberra: Coombs Academic Publishing. pp 62-68.<br />

Reeve, R., 1989. Recent Work on the Prehistory <strong>of</strong> the Western Solomons, Melanesia.<br />

Bulletin <strong>of</strong> the Indo-Pacific Prehistory Association 9: 44-67<br />

Rice, P., 1982. Pottery Production, Pottery Classification, and the Role <strong>of</strong><br />

Physicochemical Analyses. In J. S. Olin and A. D. Franklin (eds) Archaeological<br />

Ceramics. Washington DC: Smithsonian Institute Press, pp.47-56.<br />

---, 1987. Pottery Analysis: A Sourcebook. <strong>University</strong> <strong>of</strong> Chicago Press.<br />

Robertson, I. G., 2001. Mapping the Social Landscape <strong>of</strong> an Early Urban Centre. PhD<br />

thesis. Arizona State <strong>University</strong>.<br />

Roe, D., 1992. Investigations into the Prehistory <strong>of</strong> the Central Solomons: Some Old and<br />

Some New Data from Northwest Guadalcanal. In J.-C. Galipaud (ed) Poterie<br />

Lapita Et Peuplement: Actes Du Colloque Lapita, Janvier 1992. Noumea, New<br />

Caledonia: O.R.S.T.R.O.M., pp.91-101.<br />

---, 1993. Prehistory without Pots: Prehistoric Settlement and Economy <strong>of</strong> Northwest<br />

Guadalcanal, Solomon Islands. PhD thesis. Australian National <strong>University</strong>.<br />

Rogers, G., 1973. Report on Archaeological Survey in Niuatoputapu Island, Tonga.<br />

Working Papers in Anthropology, Archaeology, Linguistics and Maori Studies No.<br />

28. Department <strong>of</strong> Anthropology, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

Ross, M. D., 1996. Pottery Terms in Proto Oceanic. In J. M. Davidson, G. Irwin, B. F.<br />

Leach, A. Pawley and D. Brown (eds) Oceanic Culture History: Essays in Honour<br />

526


<strong>of</strong> Roger Green. New Zealand Journal <strong>of</strong> Archaeology Special Publication, pp.67-<br />

82.<br />

Rouse, I. B., 1960. <strong>The</strong> Classification <strong>of</strong> Artifacts in Archaeology. American Antiquity 25:<br />

313-23<br />

Rowland, M. J., 1989. Population Increase:Intensification or a Result <strong>of</strong> Preservation?<br />

Explaining Site Distribution Patterns on the Coast <strong>of</strong> Queensland. Australian<br />

Aboriginal Studies 2: 32-42<br />

Ruig, J., 1999. Collectors as Taphonomic Agents in the Archaeological Record. In M.<br />

Mountain and D. Bowdery (eds) Taphonomy: <strong>The</strong> Analysis <strong>of</strong> Processes from<br />

Phytoliths to Megafauna. Canberra: ANH publications: Australian National<br />

<strong>University</strong>, pp.109-15.<br />

Rye, O. S., 1981. Pottery Technology: Principles and Reconstruction. Manual on<br />

Archaeology 4. Washington: Taraxacum.<br />

Sand, C., 1997a. <strong>The</strong> Chronology <strong>of</strong> Lapita Ware in New Caledonia. Antiquity 71: 539-47<br />

---, 1997b. Variété De L'habitat Ancien En Nouvelle-Calédonie: Étude De Cas Sur Des<br />

Vestiges Archéologiques Du Centre-Nord De La Grande Terre. Journal, Société<br />

des Océanistes - Paris 104: 39-66<br />

---, 1998. Archaeological Report on Localities WKO013a and WKO013b at the Site <strong>of</strong><br />

Lapita(Kone, New Caledonia). Journal <strong>of</strong> the Polynesian Society 107: 7-33<br />

---, 1999. Lapita and Non-Lapita Ware During New Caledonia's First Millenium <strong>of</strong><br />

Austronesian Settlement. In J.-C. Galipaud and I. Lilley (eds) <strong>The</strong> Pacific from<br />

5000 to 2000 BP: Colonisation and Transformations. Patis: Editions de IRD,<br />

pp.139-59.<br />

---, 2000. <strong>The</strong> Specificities <strong>of</strong> the "Southern Lapita Province", the New Caledonian Case.<br />

Archaeology in Oceania 35: 20-33<br />

---, 2001. Evolutions in the Lapita Cultural Complex: A View from the Southern Lapita<br />

Province. Archaeology in Oceania 36: 65-76<br />

527


Schiffer, M. B., 1987. Formation Processes <strong>of</strong> the Archaelogical Record. Albuquerque:<br />

<strong>University</strong> <strong>of</strong> New Mexico Press.<br />

---, 1995a. Archaeological Context and Systemic Context. Behavioural Archaeology:<br />

First Principles. Salt Lake City: <strong>University</strong> Of Utah Press, pp.25-34.<br />

---, 1995b. Archaeology as Behavioural Science. Behavioural Archaeology: First<br />

Principles. Salt Lake City: <strong>University</strong> Of Utah Press, pp.46-54.<br />

---, 1995c. Toward the Identification <strong>of</strong> Formation Processes. In M. B. Schiffer (ed)<br />

Behavioural Archaeology: First Principles. Salt Lake City: <strong>University</strong> Of Utah<br />

Press, pp.171-95.<br />

Schiffer, M. B., Sullivan A. P., Klinger T. C., 1978. <strong>The</strong> Design <strong>of</strong> Archaeological<br />

Surveys. World Archaeology 10 (1): 1-28<br />

Seber, G. A. F., 1986. A Review <strong>of</strong> Estimating Animal Abundance. Biometrics 42 (2):<br />

267-92<br />

Sergeantson, S. W., Gao X., 1995. Homo Sapiens Is an Evolving Species: Origins <strong>of</strong> the<br />

Austronesians. In P. Bellwood, J. J. Fox and D. Tryon (eds) <strong>The</strong> Austronesians:<br />

Historical and Comparative Perspectives. Canberra: Department <strong>of</strong> Anthropology,<br />

Research School <strong>of</strong> Pacific and Asian Studies, Australian National <strong>University</strong>,<br />

Shackley, M. L., 1974. Stream Abrasion <strong>of</strong> Flint Implements. Nature 248: 501-2<br />

---, 1978. <strong>The</strong> Behaviour <strong>of</strong> Artefacts as Sedimentary Particles in a Fluvatile Environment.<br />

Archaeometry 20: 55-61<br />

Sharp, N. D., 1988. Style and Substance: A Reconsideration <strong>of</strong> the Lapita Decorative<br />

System. In P. V. Kirch and T. L. Hunt (eds) Archaeology <strong>of</strong> the Lapita Cultural<br />

Complex: A Critical Review. Seattle: Burke Museum, pp.61-81.<br />

---, 1991. Lapita as Text: <strong>The</strong> Meaning <strong>of</strong> Pottery in Melanesian Prehistory. Bulletin <strong>of</strong> the<br />

Indo-Pacific Prehistory Association 11: 323-32<br />

Shennan, S., 1997. Quantifying Archaeology. Edinburgh: Edinburgh <strong>University</strong> Press.<br />

Shepard, A. O., 1963. Ceramics for the Archaeologist. Washington DC: Carnegie<br />

528


Institution <strong>of</strong> Washington.<br />

Sheppard, P. J., 1993. Lapita Lithics: Trade/Exchang and Technology. A View from Thr<br />

Reefs/Santa Cruz. Archaeology in Oceania 28: 121-37<br />

---, 1996. Annual Report <strong>of</strong> the New Georgia Archaeological Survey 1996, Roviana<br />

Lagoon Survey Year 1 Centre for Archaeological Research, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong><br />

---, 2003. Personal Communication: to M. Felgate.<br />

Sheppard, P. J., Felgate M., Roga K., Keopo J., Walter R., 1999. A Ceramic Sequence<br />

from Roviana Lagoon (New Georgia, Solomon Islands). In J.-C. Galipaud and I.<br />

Lilley (eds) <strong>The</strong> Pacific from 5000 to 2000 BP: Colonization and<br />

Transformations. Paris: Editions de IRD, pp.313-22.<br />

Sheppard, P. J., Green R. C., 1991. Spatial Analysis <strong>of</strong> the Nenumbo (Se-Rf-2) Lapita<br />

Site, Solomon Islands. Archaeology in Oceania 26: 89-101<br />

Shott, M. J., 1989a. Diversity, Organization, and Behaviour in the Material Record.<br />

Current Anthropology 30 (3): 283-315<br />

---, 1989b. On Tool-Class Use Lives and the Formation <strong>of</strong> Archaeological Assemblages.<br />

American Antiquity 54 (1): 9-30<br />

---, 1996. Mortal Pots: On Use Life and Vessel Size in the Formation <strong>of</strong> Archaeological<br />

Assemblages. American Antiquity 61 (3): 463-82<br />

---, 1998. Status and Role <strong>of</strong> Formation <strong>The</strong>ory in Contemporary Archaeological Practice.<br />

Journal <strong>of</strong> Archaeological Research 6 (4): 299-329<br />

Shott, M. J., Sillitoe P., 2001. <strong>The</strong> Mortality <strong>of</strong> Things: Correlates <strong>of</strong> Use Life in Wola<br />

Material Culture Using Age-at-Census Data. Journal <strong>of</strong> Archaeological Method<br />

and <strong>The</strong>ory 8 (3): 269-302<br />

Shutler, R., Shutler M. E., 1964. Potsherds from Bougainville Island. Asian Perspectives<br />

8: 181-3<br />

Sinopoli, C. M., 1991. Approaches to Archaeological Ceramics. New York: Plenum<br />

Press.<br />

529


Siorat, J. P., 1990. A Technological Analysis <strong>of</strong> Lapita Pottery Decoration. in Lapita<br />

Design, Form and Composition: Proceedings <strong>of</strong> the Lapita Design Workshop,<br />

Canberra, Australia, December 1988, pp. 59-82. Canberra Australia: Department<br />

<strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific Studies, Australian National <strong>University</strong><br />

Skibo, J., Schiffer M. B., 1987. <strong>The</strong> Effects <strong>of</strong> Water on Processes <strong>of</strong> Ceramic Abrasion.<br />

Journal <strong>of</strong> Archaeological Science 14: 83-96<br />

Skibo, J. N., 1992. Pottery Function: A Use Alteration Perspective. Interdisciplinary<br />

Contributions to Archaeology. New York: Plenum Press.<br />

Smith, M. F., Jr., 1983. <strong>The</strong> Study <strong>of</strong> Ceramic Function from Artefact Size and Shape.<br />

PhD thesis. <strong>University</strong> <strong>of</strong> Oregon.<br />

---, 1985. Toward an Economic Interpretation <strong>of</strong> Ceramics: Relating Vessel Size and<br />

Shape to Use. In B. A. Nelson (ed) Decoding Prehistoric Ceramics. Carbondale<br />

and Edwardsville: Southern Illinois <strong>University</strong> Press, pp.254-309.<br />

Snow, B. E., Shutler R., 1985. <strong>The</strong> Archaeology <strong>of</strong> Fuga Moro Island. Cebu City: San<br />

Carlos Publications.<br />

Sokal, R. R., Sneath P. H. A., 1963. Principles <strong>of</strong> Numerical Taxonomy. San Francisco:<br />

W. H. Freeman and Co.<br />

Solow, 1994. On the Bayesian Estimation <strong>of</strong> the Number <strong>of</strong> Species in a Community.<br />

Ecology 75 (7): 2139-42<br />

Spaulding, A. C., 1953. Statistical Techniques for the Discovery <strong>of</strong> Artefact Types.<br />

American Antiquity 18: 305-14<br />

Specht, J., 1968. Preliminary Report <strong>of</strong> Excavations on Watom Island. Journal <strong>of</strong> the<br />

Polynesian Society 77: 117-34<br />

---, 1969. Prehistoric and Modern Pottery Industries <strong>of</strong> Buka Island, T.P.N.G. PhD thesis.<br />

Research School <strong>of</strong> Archaeology and Natural History, Australian National<br />

<strong>University</strong>.<br />

---, 1974. Lapita Pottery at Talasea, West New Britain. Antiquity 48: 302-6<br />

530


---, 1977. Book Review - the Lapita Pottery Style in Fiji and Its Associations. Mankind 11<br />

(1): 75-6<br />

---, 1981. Obsidian Sources at Talasea, West New Britain, Papua New Guinea. Journal<br />

<strong>of</strong> the Polynesian Society 90 (3): 337-56<br />

---, 1985. Report on Archaeological Research in the Arawe and Kandrian Areas <strong>of</strong> West<br />

New Britain Province. In J. Allen (eds) Lapita Homeland Project: Report <strong>of</strong> the<br />

1985 Field Season. Melbourne: La Trobe <strong>University</strong>, pp.48-53.<br />

---, 1991. Kreslo: A Lapita Pottery Site in Southeast New Britain, Papua New Guinea. In<br />

J. Allen and C. Gosden (eds) Report <strong>of</strong> the Lapita Homeland Project. Canberra,<br />

Australia: Department <strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific Studies, <strong>The</strong><br />

Australian National <strong>University</strong>, pp.189-204.<br />

---, 2002. Personal Communication: to M. Felgate. Noumea<br />

Specht, J., Fields J., 1984. Frank Hurley in Papua: Photographs <strong>of</strong> the 1920-1923<br />

Expeditions. Bathhurst, NSW, Australia: Robert Brown and Associates in<br />

association with the Australian Museum Trust.<br />

Specht, J., Fullagar R., Torrence R., 1991. What Was the Significance <strong>of</strong> Lapita Pottery<br />

at Talasea? Bulletin <strong>of</strong> the Indo-Pacific Prehistory Association 11: 281-94<br />

Spennemann, D. H. R., 1987. <strong>The</strong> Impact <strong>of</strong> Cyclonic Surge on Archaeological Sites in<br />

Tonga. Bulletin <strong>of</strong> the Indo-Pacific Prehistory Association 7: 75-87<br />

Spriggs, M., 1985. Archaeological Fieldwork on Nissan Island, North Solomons Province.<br />

In J. Allen (ed) Lapita Homeland Project: Report <strong>of</strong> the 1985 Field Season.<br />

Melbourne: La Trobe <strong>University</strong>, pp.58-63.<br />

---, 1990. <strong>The</strong> Changing Face <strong>of</strong> Lapita: Transformation <strong>of</strong> a Design. In M. Spriggs (ed)<br />

Lapita Design, Form and Composition: Proceedings <strong>of</strong> the Lapita Design<br />

Workshop, Canberra, Australia, December 1988. Canberra Australia: Department<br />

<strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific Studies, Australian National <strong>University</strong>,<br />

pp.83-122.<br />

531


---, 1991. Nissan, the Island in the Middle. Summary Report on Excavations at the North<br />

End <strong>of</strong> the Solomons and the South End <strong>of</strong> the Bismarcks. In J. Allen and C.<br />

Gosden (eds) Report <strong>of</strong> the Lapita Homeland Project. Canberra, Australia:<br />

Department <strong>of</strong> Prehistory, Research School <strong>of</strong> Pacific Studies, <strong>The</strong> Australian<br />

National <strong>University</strong>, pp.222-43.<br />

---, 1993. Island Melanesia: <strong>The</strong> Last 10 000 Years. In M. Spriggs, D. Yen, W. Ambrose,<br />

R. Jones, A. Thorne and A. Andrews (eds) A Community <strong>of</strong> Culture: <strong>The</strong> People<br />

and Prehistory <strong>of</strong> the Pacific. Canberra: Department <strong>of</strong> Prehistory, Research<br />

School <strong>of</strong> Pacific Studies, ANU, pp.187-205.<br />

---, 1997. <strong>The</strong> Island Melanesians. <strong>The</strong> Peoples <strong>of</strong> Southeast Asia and the Pacific.<br />

London: Blackwell.<br />

---, 2000. <strong>The</strong> Solomon Islands as Bridge and Barrier in the Settlement <strong>of</strong> the Pacific. In<br />

A. Anderson and T. Murray (eds) Australian Archaeologist: Collected Papers in<br />

Honour <strong>of</strong> Jim Allen. Canberra: Coombs Academic Publishing, ANU,<br />

---, 2001. Who Cares What Time It Is? <strong>The</strong> Importance <strong>of</strong> Chronology in Pacific<br />

Archaeology. In A. Anderson, I. Lilley and S. O'Connor (eds) Histories <strong>of</strong> Old<br />

Ages: Essays in Honour <strong>of</strong> Rhys Jones. Canberra: Pandanus Books, Research<br />

School <strong>of</strong> Pacific and Asian studies, ANU,<br />

Stewart, D. J., 1999. Formation Processes Affecting Submerged Archaeological Sites: An<br />

Overview. Geoarchaeology 14 (6): 565-87<br />

Stoddart, D. R., 1969a. Geomorphology <strong>of</strong> the Marovo Elevated Barrier Reef, New<br />

Georgia. Philosophical Transactions <strong>of</strong> the Royal Society <strong>of</strong> London, Series B,<br />

Biological Sciences 255 (800): 383-402<br />

---, 1969b. Geomorphology <strong>of</strong> the Solomon Islands Coral Reefs. Philosophical<br />

Transactions <strong>of</strong> the Royal Society <strong>of</strong> London, Series B, Biological Sciences 255<br />

(800): 355-82<br />

Summerhayes, G., 1996. Interaction in Pacific Prehistory. PhD thesis. School <strong>of</strong><br />

532


Archaeology, Faculty <strong>of</strong> Humanities, La Trobe <strong>University</strong>.<br />

---, 2000a. Lapita Interaction. Terra Australis 15. Canberra: Department <strong>of</strong> Archaeology<br />

and Natural History and Centre for Archaeological Research, Australian National<br />

<strong>University</strong>.<br />

---, 2000b. Recent Archaeological Investigations in the Bismarck Archipelago, Anir - New<br />

Ireland Province, Papua New Guinea. Bulletin <strong>of</strong> the Indo-Pacific Prehistory<br />

Association 19: 167-74<br />

---, 2000c. What's in a Pot? In A. Anderson and T. Murray (eds) Australian<br />

Archaeologist: Collected Papers in Honour <strong>of</strong> Jim Allen. Canberra: Coombs<br />

Academic Publishing, ANU, pp.291-307.<br />

---, 2001. Lapita in the Far West: Recent Developments. Archaeology in Oceania 36: 53-<br />

63<br />

---, 2002. Defining the Chronology <strong>of</strong> Lapita in the Bismarck Archipelago. In G. R. Clark,<br />

A. J. Anderson and T. Vunidilo (eds) <strong>The</strong> Archaeology <strong>of</strong> the Lapita Dispersal in<br />

Oceania: Papers from the 4th Lapita Conference, June 2000, Canberra,<br />

Australia. Pandanus Press, Research School <strong>of</strong> Pacific and Asian studies, ANU,<br />

pp.25-38.<br />

Swadling, P. L., 1976. <strong>The</strong> Occupation Sequence and Settlement Pattern on Santa Ana.<br />

In R. C. Green and M. M. Cresswell (eds) Southeast Solomons Culture History.<br />

Wellington, New Zealand: <strong>The</strong> Royal Society <strong>of</strong> New Zealand, Bulletin 11,<br />

pp.123-32.<br />

Tani, M., Longacre W. A., 1999. On Methods <strong>of</strong> Estimating Ceramic Uselife: A Revision<br />

<strong>of</strong> the Uselife Estimates <strong>of</strong> Cooking Vessels among the Kalinga, Phillipines.<br />

American Antiquity 64 (2): 299-308<br />

Tarlton, F. J., 1996. Effects <strong>of</strong> Southwest Pacific Tropical Cyclones on Archaeological<br />

Sites. MA thesis. Anthropology Department, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

Terrell, J. E., 1976. Perspectives on the Prehistory <strong>of</strong> Bougainville Island, Papua New<br />

533


Guinea: A Study in the Human Biogeography <strong>of</strong> the Southwestern Pacific. PhD<br />

thesis. Anthropology, Harvard <strong>University</strong>.<br />

Thurman, H. V., 1981. Introductory Oceanograpgy. Sydney: Charles E. Merrill Publishing<br />

Company.<br />

Torrence, R., Stevenson C., 2000. Beyond the Beach: Changing Lapita Landscapes on<br />

Garua Island, Papua New Guinea. In A. Anderson and T. Murray (eds) Australian<br />

Archaeologist: Collected Papers in Honour <strong>of</strong> Jim Allen. Canberra: Coombs<br />

Academic Publishing, ANU, pp.324-45.<br />

Torrence, R., White J. P., 2002. Tattooed Faces from Boduna Island, Papua New Guinea.<br />

In G. R. Clark, A. J. Anderson and T. Vunidilo (eds) <strong>The</strong> Archaeology <strong>of</strong> the<br />

Lapita Dispersal in Oceania: Papers from the 4th Lapita Conference, June 2000,<br />

Canberra, Australia. Pandanus Press, Research School <strong>of</strong> Pacific and Asian<br />

studies, ANU, pp.135-40.<br />

Trigger, B., 1989. A History <strong>of</strong> Archaeological Thought. Cambridge <strong>University</strong> Press.<br />

Turner, M., 2000. <strong>The</strong> Function, Design and Distribution <strong>of</strong> New Zealand Adzes. PhD<br />

thesis. Anthropology, <strong>University</strong> <strong>of</strong> <strong>Auckland</strong>.<br />

Vanderwal, R. L., 1973. <strong>The</strong> Torres Strait: Protohistory and Beyond. Occasional Papers.<br />

Anthropology Museum, <strong>University</strong> <strong>of</strong> Queensland, pp.157-94.<br />

Varien, M. D., Mills B. J., 1997. Accumulations Research: Problems and Prospects for<br />

Estimating Site Occupation Span. Journal <strong>of</strong> Archaeological Method and <strong>The</strong>ory<br />

4 (4): 141-91<br />

Varien, M. D., Potter J. M., 1997. Unpacking the Discard Equation: Simulating the<br />

Accumulation <strong>of</strong> Artefacts in the Archaeological Record. American Antiquity 62<br />

(2): 194-213<br />

Wahome, E. W., 1999. Ceramics and Prehistoric Exchange in the Admiralty Islands,<br />

Papua New Guinea. PhD thesis. Australian National <strong>University</strong>.<br />

Wandsnider, L., 1996. Describing and Comparing Archaeological Spatial Structures.<br />

534


Journal <strong>of</strong> Archaeological Method and <strong>The</strong>ory 3 (4): 319-84<br />

Ward, G. K., 1976. <strong>The</strong> Archaeology <strong>of</strong> Settlements Associated with the Chert Industry<br />

<strong>of</strong> Ulawa. In R. C. Green and M. M. Cresswell (eds) Southeast Solomons Culture<br />

History. Wellington, New Zealand: <strong>The</strong> Royal Society <strong>of</strong> New Zealand, Bulletin<br />

11, pp.161-80.<br />

Waterhouse, J. H. L., 1926. A Roviana Phrasebook. Sydney: Epworth Press.<br />

Waters, M. R., 1992. Principles <strong>of</strong> Geoarchaeology: A North American Perspective.<br />

<strong>University</strong> <strong>of</strong> Arizona Press.<br />

Weisler, M. I., 2001. Lapita Rockshelters <strong>of</strong> Elouaua and Mussau Islands (Sites EHM,<br />

EHN, EKO, EKP, and EKQ). In P. V. Kirch (ed) Lapita and Its Transformations<br />

in Near Oceania. Berkeley: Archaeological Research Facility, <strong>University</strong> <strong>of</strong><br />

California, pp.146-74.<br />

White, J. P., Coroneos C., Neall V., Boyd W., Torrence R., 2002. FEA Site, Boduna<br />

Island: Further Investigations. In S. Bedford, C. Sand and D. V. Burley (eds) Fifty<br />

Years in the Field. Essays in Honour and Celebration <strong>of</strong> Richard Shutler Jr's<br />

Archaeological Career. <strong>Auckland</strong>: New Zealand Archaeological Association<br />

Monograph 25, pp.101-8.<br />

White, J. P., Downie J. E., 1980. Excavations at Lesu, New Ireland. Asian Perspectives<br />

23 (2): 193-220<br />

White, J. P., Gorecki P., O'Brien R., Anderson P., 1985. Excavations at Bal<strong>of</strong> and<br />

Explorations Elsewhere in the New Ireland Province, 17 August - 23 September<br />

1985. In J. Allen (eds) Lapita Homeland Project: Report <strong>of</strong> the 1985 Field<br />

Season. Melbourne: La Trobe <strong>University</strong>, pp.9-.<br />

White, J. P., Harris M.-N., 1997. Changing Sources: Early Lapita Period Obsidian in the<br />

Bismarck Archipelago. Archaeology in Oceania 32: 97-107<br />

Wickler, S., 1995. Twenty-Nine Thousand Years on Buka: Long-Term Cultural Change<br />

in the Northern Solomon Islands. PhD thesis. Anthropology Department,<br />

535


<strong>University</strong> <strong>of</strong> Hawaii.<br />

---, 2001. <strong>The</strong> Prehistory <strong>of</strong> Buka: A Stepping Stone Island in the North Solomons. Terra<br />

Australis 16. Canberra: Department <strong>of</strong> Archaeology and Natural History and<br />

Centre for Archaeological Research, Australian National <strong>University</strong>.<br />

Wickler, S., Spriggs M., 1988. Pleistocene Human Occupation <strong>of</strong> the Solomon Islands,<br />

Melanesia. Antiquity 62: 703-6<br />

Willey, G. R., Phillips P., 1958. Method and <strong>The</strong>ory in American Archaeology. Chicago:<br />

<strong>University</strong> <strong>of</strong> Chicago Press.<br />

Wobst, H. M., 1983. We Cant See the Forest for the Trees: Sampling and the Shapes <strong>of</strong><br />

Archaeological Distributions. In J. A. Moore and A. S. Keene (eds)<br />

Archaeological Hammers and <strong>The</strong>ories. Academic Press, pp.37-85.<br />

Yen, D. E., 1976. Inland Settlement on Santa Cruz Island (Nendo). In R. C. Green and M.<br />

M. Cresswell (eds) Southeast Solomons Culture History. Wellington, New<br />

Zealand: <strong>The</strong> Royal Society <strong>of</strong> New Zealand, Bulletin 11, pp.203-24.<br />

536

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!