11.06.2013 Views

Bova Marina Archaeological Project - Department of Archaeology

Bova Marina Archaeological Project - Department of Archaeology

Bova Marina Archaeological Project - Department of Archaeology

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.

<strong>Bova</strong> <strong>Marina</strong><br />

<strong>Archaeological</strong><br />

<strong>Project</strong>:<br />

Survey and<br />

Excavations<br />

at<br />

Umbro<br />

Preliminary<br />

Report,<br />

1999<br />

Season<br />

edited by John<br />

Robb<br />

with contributions by <strong>Marina</strong><br />

Ciaraldi, Lin Foxhall, Doortje<br />

Van Hove, and David Yoon<br />

<strong>Department</strong> <strong>of</strong> <strong>Archaeology</strong><br />

University <strong>of</strong> Southampton<br />

Southampton SO17 1BJ<br />

United Kingdom<br />

tel. 00-44-23-80592247<br />

fax 00-44-23-80593032<br />

email jer@soton.ac.uk


ACKNOWLEDGEMENTS<br />

2<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

I would like to thank Dottoressa Elena Lattanzi, Soprintendente, and Dottoressa Emilia Andronico,<br />

Ispettrice, <strong>of</strong> the Soprintendenza Archeologica della Calabria, for their help and guidance in this research. As in<br />

past years, we are grateful to Sebastiano Stranges and to Luigi Saccà for their help, advice, and friendship; to<br />

Brian McConnell and Laura Maniscalco for archaeological advice; to our landlords, Antonino and Silvana<br />

Scordo, and to our cooks, Mariella Catalano and Annunziata Caracciolo. Mary Anne Tafuri translated the<br />

project summary and Fiona Coward helped assemble the final documentation. Finally, I would like to thank the<br />

field staff (David Yoon, Lin Foxhall, Paula Lazrus, Keri Brown and Starr Farr) and post-excavation staff and<br />

analysts (Umberto Albarella, <strong>Marina</strong> Ciaraldi, Sonia Collins, Kathryn Knowles, Doortje Van Hove, Jayne Watts,<br />

and David Williams) for their expertise, and all the crew members listed below for their hard work and<br />

enthusiasm.<br />

We gratefully acknowledge funding from the British Academy (Excavation Grant A-AN4798/APN7493<br />

Supplementary post-excavation funding), the Arts and Humanities Research Board (Research Grant AHRB/RG-<br />

AN4798/APN8592), the <strong>Department</strong> <strong>of</strong> <strong>Archaeology</strong>, University <strong>of</strong> Southampton, and the School <strong>of</strong><br />

<strong>Archaeology</strong> and Ancient History, University <strong>of</strong> Leicester.<br />

BOVA MARINA ARCHAEOLOGICAL PROJECT: CREW, 1999 SEASON<br />

Director John Robb<br />

Survey Co-Director David Yoon<br />

Survey Co-Director Lin Foxhall<br />

Field Supervisor Keri Brown<br />

Field Supervisor Paula Kay Lazrus<br />

Lab Manager Starr Farr<br />

Artist (Southampton) Kathryn Knowles<br />

Computing (Southampton) Doortje Van Hove<br />

Faunal analysis (Birmingham) Umberto Albarella<br />

Paleobotany (Birmingham) <strong>Marina</strong> Ciaraldi<br />

Cook Mariella Catalano<br />

Cook Annunziata Caracciolo<br />

Crew Members Siân Anthony<br />

Francesca Binyon<br />

Rebecca Crowson-Towers<br />

Glenn Dunaway<br />

Lauren Dumford<br />

Mark Ellis<br />

Anne Forbes<br />

Helen Forbes<br />

Janet Forbes<br />

Lucy Heaver<br />

Alex Hopson<br />

Jenny House<br />

Charlotte John<br />

Claire Rees<br />

Kathryn Simms<br />

Barney Skinner<br />

Jayne Watts<br />

Matthew Wortley<br />

Child Johanna Farr<br />

Infant Nicholas Robb


RIASSUNTO ITALIANO<br />

3<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

Nel 1999, il Progetto Archeologico <strong>Bova</strong> <strong>Marina</strong> ha intrapreso la sua terza campagna di scavo. Le<br />

finalità principali della ricerca includevano l’attività di ricognizione di varie zone intorno al comune di <strong>Bova</strong><br />

<strong>Marina</strong> e lo scavo del sito preistorico di Umbro. Tali attività hanno avuto luogo dal 29 agosto al 26 settembre<br />

1999, con un gruppo di 25 persone. I responsabili dello scavo erano John Robb (scavo preistorico e<br />

amministrazione generale), David Yoon (ricognizione di superficie, archeologia Romana) e Lin Foxhall<br />

(ricognizione di superficie, archeologia Greca).<br />

Ricognizione di superficie<br />

Finalità e metodi. Lo scopo principale dell’attività di ricognizione era rivolto alla comprensione dello<br />

sviluppo, avvenuto nel corso della preistoria ed in epoca storica, dei modelli insediamentali legati al territorio di<br />

<strong>Bova</strong> <strong>Marina</strong>. Si è continuato quanto intrapreso nel 1997 e 1998. Il metodo utilizzato consiste nel systematic<br />

transect walking, mediante il quale gruppi di 4-5 persone, poste ad intervalli di 10 m., camminano su territori<br />

delimitati registrandone le caratteristiche storiche e geografiche. Nel 1999, il nostro gruppo di ricognitori ha<br />

coperto una superficie di 102 ettari, portanto l’area totale finora esplorata sistematicamente a 274 ettari. Le zone<br />

ricognite comprendono due grossi appezzamenti contigui (intorno ad Umbro e nella zona costiera di San<br />

Pasquale), e numerose porzioni di territorio sparse intorno al comune. I lavori del 1997 e 1998 si erano<br />

concentrati prevalentemente nei territori di Umbro e San Pasquale. Nel 1999, la direzione della ricerca era<br />

rivolta ad estendere la porzione di territorio ispezionato ad ambienti nuovi come colline e pianori interni.<br />

Tutti i ritrovamenti sono stati catalogati e datati per quanto possibile. E’ da tenere presente tuttavia che<br />

la sequenza cronologia della ceramica presente in questa regione è piuttosto variegata. Alcuni siti preistorici,<br />

chiaramente Neolitici, possono essere facilmente individuati grazie alle decorazioni presenti sui frammenti<br />

ceramici. Tuttavia altri periodi, quali l’Eneolitico e l’età del Bronzo, sono caratterizzati da ceramica inornata la<br />

cui identificazione si basa, per la maggior parte dei casi, sulla forma del vaso, rendendo quindi molto difficile la<br />

lettura di frammenti di piccole dimensioni. Per il periodo storico, la ceramica Greca e Romana è facilmente<br />

identificabile, tuttavia, pochi dati si posseggono sulla ceramica medievale di questa regione, forse a causa dello<br />

scarso numero di siti medievali individuati.<br />

Risultati. Fino ad oggi sono stati identificati 37 siti che variano da una piccola concentrazione di<br />

frammenti preistorici non databili a grandi villagi di epoca Greca e Romana con complessa articolazione interna.<br />

I siti sono elencati e descritti al paragrafo 2.1.4. E’ da tenere presente che tale elenco aggiorna quello<br />

precedenemente fornito in quanto alcuni siti scoperti nel passato sono stati successivamente ridatati.<br />

Nel considerare i siti in ordine cronologico, fino ad oggi, non si ha traccia di attività Paleolitica o<br />

Mesolitica. I siti di epoca Neolitica sono manifestati in una numerosa varietà che comprende terrazzi costieri,<br />

basse colline prospicenti valli fluviali e pianori interni. Non si sono evidenziati insediamenti riferibili<br />

all’Eneolitico o all’età del Bronzo (benché tracce di frequentazione di età del Bronzo sono state messe in<br />

evidenza durante gli scavi ad Umbro, cfr. più avanti); tale aspetto può essere riferibile in parte a problemi di<br />

datazione dei pezzi raccolti durante l’attività di ricognizione (vedi sopra). Le testimonianze di epoca Greca<br />

includono un grande villaggio (Mazza, cfr. più avanti), e numerosi piccoli cascinali interni (siti 18 e 33). Molti<br />

sono situati su alti terrazzi. In epoca Romana si hanno per la prima volta tracce di siti su fondo valle, che<br />

suggeriscono il passaggio a forme di agricoltura intensiva. I resti ceramici di epoca Greca e Romana mostrano<br />

due apici: uno che va dal periodo Tardo Arcaico a quello Classico (VI-V secolo a.C.) e l’altro in epoca<br />

Imperiale (III-V sec d.C.). Gli insediamenti più tardi sembrano occupare gli stessi territori, ma in minore<br />

consistenza, fino all’VIII sec. circa. L’epoca Medievale rimane oscura tuttavia nel XIX sec. si registrano di<br />

nuovo numerosi insediamenti a carattere rurale.<br />

Oltre all’attività di ricognizione sono state portate a termine numerose indagini secondarie rivolte allo<br />

studio dell’uso del territorio. Esse comprendono il riconoscimento, in ambito geologico, di quelle materie prime<br />

che possono essere state utili in epoca preistorica (da continuare nel 2000), e il GIS (Geographical Information<br />

System) computer modelling volto alla ricostruzione delle antiche dinamiche di sfruttamento del territorio in<br />

termini di esigenze relative ai diversi modelli economici preistorici.<br />

Raccolta di superficie intensiva dal sito di Mazza. Si è intrapresa una raccolta di superficie intensiva<br />

sul sito Greco di Mazza. Questo ampio villaggio ha finora restituito le uniche ceramice Greche di imporazione<br />

del periodo arcaico/classico, un possibile frammento Attico ed un possibile frammento Laconico. Le finalità<br />

della raccolta intensiva erano di determinare i limiti geografici, l’intervallo cronologico e la distribuzione<br />

spaziale interna del sito. Come prima cosa è stata stabilita una griglia di punti di riferimento su tutto il sito e<br />

sono stati collocati punti di raccolta ogni 30 metri all’interno della stessa. Si è poi proceduto alla raccolta di tutti<br />

i manufatti presenti in un’area di 10 metri quadrati per ogni punto di raccolta. I risultati potranno essere


4<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

analizzati quantitativamente. L’analisi dei dati provenienti dal sito di Mazza è ancora in corso, tuttavia alcune<br />

considerazioni sono già possibili. Il sito sembra possedere una componente Romana preponderante che si<br />

aggiunge alle evidenze di epoca Greca. Un considerevole numero di tegole di epoca Romana, insieme a tegole di<br />

epoca Greca, è stato ritrovato nella parte alta del sito. Tegole di epoca Greca sono state ritrovate in piccole<br />

quantità sul resto della zona superiore del sito e sul fianco sud-est della collina. La ceramica fine Greca<br />

comprende una considerevole quantità di tazze di fattura locale e forme a cratere. Gli affioramenti rocciosi<br />

presenti nella parte centrale del sito somigliano a quelli utilizzati nelle fondazioni di Locri Epizephyrii, e<br />

possono essere aver avuto lo stesso scopo. L’insediamento si estende con minore densità su tutto il pianoro di<br />

Mazza. Il limite meridionale del sito potrebbe essere stato, in antico, una zona industriale, come testimonia l’alta<br />

concentrazione di resti di fusione e lavorazione dei metalli. Resti di concotto e scorie di fornace sono stati<br />

raccolti e sono attualmente in corso di analisi presso l’Univesità di Leicester.<br />

Il sito di Mazza è stato abitato lungo un ampio arco cronologico durante il periodo classico. I materiali<br />

raccolti hanno indicato la presenza sia di abitazioni che di attività industrali nonché di un probabile grande<br />

edificio in prossimità della sommità del sito. Si è accertata inoltre la presenza di ceramica preistorica. Studi<br />

futuri sul sito si concentreranno sulla relazione tra lo stesso e le città di Reggio e Locri Epizephyrii e sulle<br />

problematiche relative all’occupazione indigena della zona. L’individuazione, nell’area studiata, di una serie di<br />

piccoli insediamenti rurali Greci e la considerevole distanza da qualsiasi città Greca rappresentano un<br />

interessante aspetto da analizzare nella comprensione della colonizzazione Greca e dell’occupazione di territori<br />

rurali a scopo coloniale. In genere, gli studi sulla colonizzazione Greca si sono concentrati su informazioni<br />

provenienti da siti urbani come Locri e dalle zone circostanti come Metaponto. Le datazioni antiche di alcuni siti<br />

provenienti dalla ricognizione di <strong>Bova</strong> sono sorprendenti, esse potrebbero indicare che l’insediamento Greco si<br />

sia diffuso ampiamente sul territorio coloniale nell’arco di poche generazioni, o che le popolazioni indigene si<br />

siano, almeno in parte, ellenizzate piuttosto rapidamente. La funzione di questi siti, e la loro relazione con le<br />

città Greche rimane piuttosto incerta, non si hanno inoltre informazioni circa una occupazione autoctona che sia<br />

essa contemporanea o leggermente più antica. Lo scavo di un insediamento Greco di piccole dimensioni come il<br />

sito 18 (Umbro) potrebbe fornire dati da mettere in relazione con siti urbani come Locri o ‘cascinali’ presenti su<br />

territorio Greco, allo scopo di risolvere alcune questioni.<br />

Note su siti soggetti a minaccia. Benché numerosi degli insediamenti presenti nel comune di <strong>Bova</strong><br />

<strong>Marina</strong> siano stati danneggiati dall’erosione o da processi geomorfologici, alcuni dei siti osservati sono risultati<br />

soggetti ad azioni distruttive provocate dall’uomo quali, ad esempio, l’attività edilizia. Tutti i siti minacciati sono<br />

a conoscenza della Soprintendenza; questi includono principalmente la villa Romana o cascinale di Panaghia,<br />

San Pasquale, in corso di danneggiamento da attività di costruzione e attualmente vincolato, e la villa Romana o<br />

insediamento urbano di Torrente Siderone (Amigdala), sotto la SS 106 presso la città moderna di <strong>Bova</strong> <strong>Marina</strong>,<br />

attualmente in corso di scavo da parte della Soprintendenza.<br />

Scavi ad Umbro<br />

Umbro appare come un complesso sito a più fasi, posto a circa 4 km nell’entroterra, in prossimità<br />

dell’antico confine tra i comuni di <strong>Bova</strong> <strong>Marina</strong> e <strong>Bova</strong> Superiore. Scoperto nel 1990 da S. Stranges e L. Saccà,<br />

il sito è stato soggetto a ripetute attività di raccolta di superficie e, nel 1998, ad una nostra prima campagna di<br />

scavo. Lo scavo del 1998 stabilì che l’insediamento consisteva di una piccola area, alla base di un dirupo<br />

intensamente occupato durante il Neolitico, e di numerose concentrazioni di materiale, alla sommità dello stesso,<br />

probabilmente di età del Bronzo. L’insediamento Neolitico principale ha restituito una datazione al<br />

radiocarbonio che va dalla metà del VI millennio BC alla metà del V millennio BC (Stentinello). E’ inoltre<br />

attestata l’occupazione durante la facies di Diana e, probabilmente in maniera sporadica, durante l’età del Rame.<br />

Nel 1999, i lavori si sono concentrati in tre trincee: la Trincea 1, la Trincea 6 e la Trincea 7.<br />

Trincea 1: l’area Neolitica. Tale trincea fu iniziata sott<strong>of</strong>orma di sondaggio stratigrafico nel 1998 ed<br />

esteso allo scopo di comprendere la zona centrale del sito di epoca Neolitica posto alla base del dirupo. Essa<br />

comprendeva una trincea di un metro per sei metri, con una piccola estensione laterale al centro. Gli scavi sono<br />

proceduti in aree di un metro quadrato con tagli arbitrari di 10 cm.; il terreno rimosso è stato setacciato in griglie<br />

di 5 mm. Gli scavi sono proceduti fino ad arrivare alla base rocciosa nella parte meridionale della trincea; nel<br />

margine settentrionale della stessa lo scavo si è interrotto ad 1,5 m. dal piano di calpestio; il materiale antropico<br />

presente a questa pr<strong>of</strong>ondità era considerevolmente diminuito.<br />

La Trincea 1 ha restituito numerosi manufatti Neolitici, ma nessuna evidenza di strutture abitative quali<br />

capanne. E’ evidente che il sito consisteva in un’area ristretta (con uno spazio abitativo di forse 20 metri<br />

quadrati), occupata da piccoli gruppi di persone lungo il corso di diversi millenni. Tutti i manufatti erano<br />

considerevolmente frammentati. Sono stati identificati in tutto cinque strati. Lo Strato I consisteva di 5-10 cm. di<br />

terreno di superficie inquinato. Lo Strato II era formato da uno spesso livello di sedimento bruno chiaro con


5<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

abbondante presenza di materiale roccioso di crollo, probabilmente disturbato sian in epoca recente che in<br />

antico. Gli Strati I e II contenevano manufatti di tutti i periodi. Lo Strato III, di sedimento chiaro giallognolo,<br />

conteneva materiale roccioso di crollo di diverse dimensioni e una serie di lenti sabbiose apparentemente<br />

derivate dal disgregamento dei frammenti rocciosi. Il materiale ceramico presente è riferibile alle facies di<br />

Stentinello e Diana, e sembra essere riferibile a fasi di Diana con presenza di intrusioni di epoche precedenti; ciò<br />

non sorprende visto che l’area Neolitica è posta sul fondo di un bacino roccioso. Lo Strato IV consiste di un<br />

terreno più denso e compatto di colore bruno. Esso contiene prevalentemente ceramica di tipo Stentinello ed è<br />

databile al V millennio BC. In ultimo, lo Strato V è alla base dell’area scavata e contiene terreno argilloso di<br />

colore bruno scruro con presenza di ceramica Stentinello databile al VI millennio BC. I manufatti Neolitici<br />

raccolti comprendono ceramica tipica degli stili di Stentinello e Diana, alcuni esempi di ceramia Neolitica di<br />

altro tipo, quali ceramica a pittura rossa, ed alcuni probabili frammenti di epoca Eneolitica (prevalentemente<br />

dagli Strati II e III). L’industria litica era ricca di ossidiana ed includeva numerose piccole schegge e lamelle.<br />

Sono state ritrovate un’ascia in anfibolite ed una piccola riproduzione di ascia in fillite, si è attestata inoltre la<br />

presenza di numerosi ciottoli in pietra metamorfica di importazione, alcune macine e alcuni frammenti di<br />

concotto e di ocra rossa. I resti faunistici non sono ancora stati studiati, tuttavia è già evidente la preponderanza<br />

degli ovini e dei caprini, con presenza di maiale; i bovini sono scarsamente rappesentati e l’itti<strong>of</strong>auna è assente.<br />

Sono stati conservati campioni di flottazione provenienti da ogni Strato. L’analisi dei semi e dei resti vegetali<br />

carbonizzati mostrano che la coltura di alcune specie era praticata; i resti vegetali includono la veccia, il<br />

Triticum aestivum s.l., il Triticum dicoccum, e l’orzo.<br />

La Trincea 6: l’area di età del Bronzo. Nel 1999 si iniziata la pulitura del pr<strong>of</strong>ilo della rupe dal quale la<br />

ceramica dilavava dal bordo accanto alla strada che attraversa il sito. Dal fianco del dirupo è emerso un vaso<br />

intero, si è così deciso di aprire una trincea per effettuare ulteriori indagini. Nel complesso si è messa in<br />

evidenza la deposizione di un gruppo di tre vasi comprendenti due tazzine attingitoio con ansa soprelevata ed<br />

una grande olla dotata di tre anse orizzontali poste vicino all’orlo. I vasi erano affiancati e posti su uno stesso<br />

piano di calpestio distinguibile da quella di riempimento sopra per la grandezza e l’orientamento delle pietre. Si<br />

è attestata la presenza di piccoli frammenti di concotto e carbone. Subito sotto il gruppo di vasi è stato ritrovato<br />

un inusuale frammento di corno fittile. Sulla superficie del battuto si è attestata la presenza di un crollo con<br />

pietre di medie e grandi dimensioni (fino a 30 cm. ca.). Il crollo sembrava di maggiore volume ed intensità ad est<br />

del gruppo di vasi. Non è chiaramente comprensibile se ci si trovi di fronte ad un fenomeno erosivo naturale o al<br />

crollo di qualche strutture abitativa. In termini di datazione, i vasi sono chiaramente riferibile al Bronzo Antico<br />

ed il corno fittile è tipico dello stile Rodì-Tindari-Vallelunga. In stretto accordo con quanto suggerito dallo stile<br />

ceramico, la datazione al radiocarbonio, fornita da un frammento di carbone proveniente dal piano di posa del<br />

gruppo dei vasi, è riferibile al 1720-1580 BC (cal.)(intervallo 2-sigma).<br />

Il gruppo ceramico della Trincea 6 potrebbe rappresentare quanto rimane del crollo di una capanna,<br />

benchè sono scarsi i resti di frequentazione domestica all’interno della trincea e segni di strutture tipiche dell’età<br />

del Bronzo quali capanne con soglie in pietra. Una seconda possibilità potrebbe essere rituale: i vasi potrebbero<br />

rappresentare una qualche deposizione rituale idiosincratica. Una terza possibilità è rappresentata da una<br />

eventuale deposizione funeraria. Confronti effettuati con contesti della Sicilia orientale hanno messo in evidenza<br />

la presenza ricorrente, durante il Bronzo Antico, di sepolture situate lungo margini scoscesi di pietra calcarea o<br />

di altro tipo, così come appare essere Umbro. Si tratta in genere di piccole camere artificiali raramente più<br />

grandi dei di 2 metri di diametro e 1,5 metri di altezza, spesso raggiungibili attraverso uno stretto passaggio. In<br />

diversi casi, inoltre, gli archeologi hanno messo in luce aree livellate o piattaforme che contenevano gruppi di<br />

vasi rappresentanti il corredo funerario volontariamente posto all’esterno della tomba piuttosto che all’interno di<br />

essa (es. Santa Febronia, Maniscalco, 1996).<br />

Trincea 7. La Trincea 7 consiste di un piccolo (un metro per un metro) pozzetto scavato<br />

immediatamento al di sotto del margine del dirupo, lungo il limite meridionale del bacino che racchiude il sito<br />

Neolitico. Si è sperato di recuperare in questa area materiale di epoca preistorica. Tuttavia, pochi resti sono stati<br />

ritrovati, e ad una pr<strong>of</strong>ondità di 1,5 metri lo scavo della trincea è risultato impraticabile.<br />

Sondaggi di Scavo a San Pasquale<br />

Nel 1999 è stato sondato il sito Neolitico di San Pasqule, situato su di un basso terrazzo lungo il<br />

margine orientale della valle di S. Pasquale. Si è sperato di verificare la presenza di un “villaggio” Neolitico, per<br />

il quale la zona <strong>of</strong>friva le caratteritiche tipiche. Due pozzetti di un metro per un metro sono stati scavati in una<br />

zona situata lungo il margine occidentale del sito, dove in superficie emergevano frammenti di ceramica<br />

preistorica. I risultati si sono rivelati piuttosto scoraggianti. Entrambi i sondaggi hanno evidenziato che non<br />

esistevano depositi archeologici al di sotto dello strato superficiale; il terreno si è rapidamente trasformato in uno<br />

strato di sabbia sterile. Appare evidente che il materiale ceramico di epoca preistorica proviente dal “sito” deriva


6<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

dal dilavamento della collina posta a nord-est di esso, in parte coltivata ed in parte coperta da una fitta macchia<br />

mediterranea. Indagini future potrebbero concentrarsi sull’individuazione del sito originario.<br />

Indagini Future<br />

La campagna di scavo del 1999 ha aperto la strada a nuove prospettive di ricerca.<br />

Ricognizione. Nelle campagne future, speriamo di poter continuare ed incrementare l’attività di<br />

ricognizione, che stà fornendo nuove informazioni sull’utilizzazione del territorio. Allo stesso tempo speriamo di<br />

indagare due problemi metodologici.<br />

Ricognizioni di altura a <strong>Bova</strong> Superiore. Sia le indagini storiche che la nostra ricostruzione informatica<br />

al GIS hanno dimostrato l’importanza, in diversi periodi, degli insediamenti d’altura in zone interne. Un<br />

sondaggio basato solo sull’indagine di zone costiere sarebbe inevitabilmente incompleto e non potrebbe fornire<br />

l’intera consistenza e l’evoluzione delle dinamiche insediamentali del territorio di <strong>Bova</strong> <strong>Marina</strong>; questo potrebbe<br />

spiegare inoltre la scarsa presenza di siti di età del Ferro o di epoca Medievale. Nel continuare i sondaggi a <strong>Bova</strong><br />

<strong>Marina</strong>, speriamo inoltre di poterci estendere al comune di <strong>Bova</strong> Superiore e di completare il quadro delle<br />

conoscenze sulle strategie insediamentali antiche, investigando un completa serie di diversi ambienti.<br />

Sondaggi di scavo in siti di superficie non datati. Sono inoltre stati ritrovati una serie di siti non datati<br />

che potrebbero rappresentare i nosti “periodi mancanti”. L’unico modo per indagare gli stessi consiste nello<br />

scavo programmato di sondaggi di un metro quadrato, in grado di fornire materiale datante meno danneggiato,<br />

campioni per datazioni assolute al radiocarbonio, ed informazioni sullo stato di conservazione del sito.<br />

Nell’immediato futuro ci auguriamo di poter sondare due aree entro un raggio di 200-300 metri dallo scavo di<br />

Umbro: Limaca, dove una concentrazione di materiale potrebbe rappresentare una successiva occupazione di età<br />

del Bronzo (si è ottenuto un permesso nel 1999 che è però arrivato troppo tardi perché si potesse praticamente<br />

procedere allo scavo); e una zona a circa 200 metri a sud-ovest di Umbro dove due concentrazioni di frammenti<br />

di epoca preistorica sono visibili in superficie. Quest’ultima potrebbe rappresentare un villaggio Neolitico<br />

all’aperto; si considera critica una eventuale verifica che possa permettere di comprendere le dinamiche<br />

insediamentali del territorio di Umbro.<br />

Lo scavo ad Umbro: Ci auguriamo di poter continuare gli scavi sia del sito Neolitico che di quello di<br />

età del Bronzo dell’area di Umbro. Gli scavi saranno limitati nelle finalità e si concentreranno sulla<br />

individuazione della base della stratigrafia, sulla verifica dell’esistenza di eventuali frequentazioni pre-<br />

Neolitiche e sul recupero di un maggior numero di campioni per analisi, per esempio, di tipo paleobotanico.<br />

Nell’area di età del Bronzo Antico (Trincea 6), speriamo di poter verificare la natura del gruppo ceramico,<br />

ovvero se formi o meno parte di una sepoltura.<br />

Altre attività di scavo: Come già accennato, l’area di Umbro è ricca di piccole concentrazioni di<br />

materiale archeologico. In alcune di esse ci auguriamo di poter effettuare dei sondaggi esplorativi (vedi sopra).<br />

Un ulteriore sito da indagare è rappresentato dal piccolo insediamento Greco posto a circa 200 meri a sud di<br />

Umbro. La presenza di piccoli insediamenti interni di epoca greca, come il cascinale di Umbro, si sono rivelati<br />

una scoperta sorprendente e potrebbero risultare di notevole importanza nella comprensione del processo di<br />

colonizzazione di questa zona.<br />

Note sull’organizzazione della documentazione<br />

La documentazione relativa alla campagna del 1999 risulta divisa in due parti. La prima parte (questo<br />

volume) consiste in un rapporto descrittivo illustrato. La seconda parte è formata dalla compilazione di dati di<br />

appendice per il progetto e per gli archivi della Soprintendenza. Sono inclusi cataloghi degli oggetti suddivisi<br />

per provenienza e categoria di oggetto, disegni di tutti i frammenti diagnostici, il giornale di scavo e schede<br />

standard per siti, unità stratigrafiche e frammenti ceramici.


CONTENTS<br />

7<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

Riassunto Italiano 3<br />

1. Introduction 8<br />

2. Field Survey and Landscape Studies 9<br />

2.1. Field Survey (David Yoon) 9<br />

2.2. Intensive collections at Mazza (Lin Foxhall) 19<br />

2.3. Geological reconnaissance 20<br />

2.4. GIS analysis <strong>of</strong> land use (Doortje Van Hove) 21<br />

3. The Excavations at Umbro 24<br />

3.1. Introduction: previous work, goals and methods for this season 24<br />

3.2. Description <strong>of</strong> 1999 Trenches 25<br />

3.3. Other areas explored 30<br />

4. Umbro Excavations: Description <strong>of</strong> Finds 32<br />

4.1. Ceramics 32<br />

4.2. Lithics 34<br />

4.3. Other Artifacts: daub, worked shell and bone, ground and polished stone, and ochre 35<br />

4.4. Faunal Remains and Human Remains 37<br />

4.5. Paleobotanical Remains (<strong>Marina</strong> Ciaraldi) 38<br />

5. Test Excavations at San Pasquale 41<br />

6. Conclusions and Future Research Directions 43<br />

6.1. Umbro: dating, stratigraphy, and site function 43<br />

6.2. The Landscape <strong>of</strong> <strong>Bova</strong> <strong>Marina</strong> from Neolithic through Recent Times 43<br />

6.3. Future Work 43<br />

Bibliography 45<br />

Figures and Tables 47<br />

NOTE ON THE ORGANIZATION OF THIS REPORT<br />

Description <strong>of</strong> the 1999 field season is presented in two parts. This report, which comprises the first part,<br />

presents a self-contained, synthetic description <strong>of</strong> the survey, excavations and finds. The second part presents<br />

detailed records <strong>of</strong> various kinds for archive purposes: catalogs <strong>of</strong> finds inventoried within each artifact bag and<br />

listed by kind <strong>of</strong> artifact, ceramic inventories and drawings <strong>of</strong> all diagnostic sherds, the excavation diary, and so<br />

on.


1999 was the third year <strong>of</strong> the <strong>Bova</strong><br />

<strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong>. In 1997, a small<br />

crew <strong>of</strong> 5 Southampton students led by John Robb<br />

carried out field survey in the comune <strong>of</strong> <strong>Bova</strong><br />

<strong>Marina</strong>. In 1998, a larger crew <strong>of</strong> 10 Southampton<br />

students and about 5 staff, directed by John Robb<br />

and David Yoon, surveyed further areas, and<br />

excavated numerous exploratory trenches at the<br />

Neolithic site <strong>of</strong> Umbro.<br />

These campaigns are described in the<br />

1997 and 1998 preliminary reports (Robb 1997;<br />

1998), which also detail the circumstances in which<br />

fieldwork was carried out and the geography,<br />

geology, and archaeology <strong>of</strong> the region. To avoid<br />

repetition, these topics will not be extensively<br />

described here.<br />

In 1999, we had the largest field season to<br />

date, with 18 (10 from Southampton and 8 from<br />

Leicester) crew members and 10 staff or<br />

contributing specialists, directed by John Robb,<br />

David Yoon and Lin Foxhall. We worked in <strong>Bova</strong><br />

<strong>Marina</strong> from August 29 through September 25. The<br />

season had three major goals:<br />

(1) to carry out extensive excavations in<br />

the Neolithic area <strong>of</strong> Umbro, with particular<br />

attention to economic and contextual information<br />

about the Neolithic habitation there;<br />

1. INTRODUCTION<br />

8<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

(2) to carry out test excavations in the<br />

Neolithic site <strong>of</strong> San Pasquale, in order to assess its<br />

potential for large-scale excavations;<br />

(3) to continue the field survey, extending<br />

it to new areas, particularly inland; the overall<br />

survey goal is to understand the evolution <strong>of</strong><br />

settlement in the area <strong>of</strong> <strong>Bova</strong> <strong>Marina</strong> throughout<br />

both prehistoric and historic times.<br />

In addition, we aimed to carry out a number <strong>of</strong><br />

complementary minor projects such as GIS<br />

environmental modelling, geological raw materials<br />

survey, and intensive gridded collection at the<br />

Greek site <strong>of</strong> Mazza with the goal <strong>of</strong> clarifying its<br />

internal organization and evolution.<br />

The 1999 field season ran smoothly, with<br />

no major hitches, and these goals were<br />

accomplished. In some cases the results were<br />

disappointing, as in the San Pasquale test<br />

excavations, but in other cases surprises emerged,<br />

as in the Bronze Age ritual deposition discovered at<br />

Umbro. The purpose <strong>of</strong> this report is to describe the<br />

methods and results <strong>of</strong> this field season in detail,<br />

and to interpret their significance for both the<br />

archaeology <strong>of</strong> the region and the development <strong>of</strong><br />

future fieldwork.


2.1. Field Survey (David Yoon)<br />

2.1.1. Background<br />

The overall goal <strong>of</strong> the BMAP field survey<br />

is to understand how people used different areas <strong>of</strong><br />

the Calabrian landscape for various purposes such<br />

as habitation, farming, herding, foraging,<br />

specialized production, and so on, in all periods<br />

from the Paleolithic through modern times. As<br />

detailed in previous reports, the local landscape is<br />

both very rugged, with very little level ground, and<br />

highly varied, with very diverse environments<br />

located close to each other. The contrast between<br />

narrow coastal strips and valley bottoms and<br />

mountainous interior is especially marked.<br />

Historically, there appear to have been oscillations<br />

between coast-oriented settlement such as in<br />

Roman times and inland settlement as in medieval<br />

times. As also detailed previously, a number <strong>of</strong><br />

landscape features make archaeological field survey<br />

challenging and its results sometimes problematic.<br />

These include heavy erosion on slopes and deep<br />

alluviation on valley bottoms, intensive and<br />

destructive historical land use, the rapid<br />

proliferation <strong>of</strong> fenced-in non-surveyable areas, and<br />

the fact that pottery fragments from many periods<br />

are not particularly diagnostic.<br />

In 1997, a strategy <strong>of</strong> visiting previously<br />

known sites and systematically walking a variety <strong>of</strong><br />

small areas provided an efficient introduction to the<br />

local archaeological sequence and also provided<br />

useful experience for adapting survey methods to<br />

local conditions. In 1998, the same survey strategy<br />

was continued, together with an effort to assemble<br />

larger contiguous tracts <strong>of</strong> systematically surveyed<br />

territory, especially around Umbro and San<br />

Pasquale.<br />

The work in 1997 and 1998 raised a<br />

number <strong>of</strong> questions which we hoped to address in<br />

1999. Some periods (Paleolithic, Copper Age, Iron<br />

Age, Medieval) were absent or nearly absent from<br />

our collections. Does this represent an actual<br />

scarcity <strong>of</strong> occupation in those periods or did it<br />

instead reflect poor recognition <strong>of</strong> artifacts by us or<br />

poor choice <strong>of</strong> survey areas? The recognition <strong>of</strong><br />

several small Greek sites <strong>of</strong> early date raised many<br />

questions about the process <strong>of</strong> Greek colonization<br />

in the region, and how pre-existing native<br />

communities interacted with the Greek presence.<br />

The differences observed so far in Greek and<br />

Roman site locations (Greek sites on hills, Roman<br />

sites in valleys), despite similarities <strong>of</strong> technology<br />

and social organization, should be associated with<br />

2. FIELD SURVEY AND LANDSCAPE STUDIES<br />

9<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

important differences in land use and economic<br />

organization.<br />

In the 1999 survey, co-directed by David<br />

Yoon and Lin Foxhall, we therefore sought in 1999<br />

to improve the quality <strong>of</strong> our evidence as it relates<br />

to these questions, by:<br />

• doing more survey <strong>of</strong> hilltops and highelevation<br />

areas<br />

• continuing to build up larger blocks <strong>of</strong><br />

completely surveyed territory<br />

• obtaining more detailed information about the<br />

locations and internal organization <strong>of</strong> Greek<br />

and Roman sites<br />

Survey <strong>of</strong> hilltops and high-elevation areas was<br />

intended to improve the representation <strong>of</strong> these in<br />

our sample, and particularly to see whether the<br />

missing time periods would be located in these<br />

places. Larger contiguous blocks <strong>of</strong> surveyed<br />

territory enable better understanding <strong>of</strong> the relative<br />

placement <strong>of</strong> contemporary sites, which is<br />

important for interpreting the organization <strong>of</strong> local<br />

communities <strong>of</strong> any period. Because Greek and<br />

Roman village sites are relatively large and<br />

complex, they are difficult to interpret through<br />

ordinary fieldwalking methods. The controlled<br />

surface collections at Deri (Site 9, San Pasquale) in<br />

1997 provided some information on the internal<br />

structure <strong>of</strong> one large Roman site. We decided in<br />

1999 to do similar controlled surface collections at<br />

Mazza, the largest known Greek site in the study<br />

area, to get comparable information for that period.<br />

2.1.2. Procedures<br />

The methods used in 1999 were essentially<br />

the same as in previous years. We worked in crews<br />

consisting either <strong>of</strong> one <strong>of</strong> the co-directors and<br />

three to four students or (more rarely) both codirectors<br />

and six to eight students. All survey was<br />

carried out within defined “areas,” zones .1-2 ha in<br />

size whose boundaries, location, and geology and<br />

land use were recorded systematically, All areas<br />

without previously identified sites were walked in<br />

parallel transects with the crew members spaced 10<br />

meters apart (in some locations, where it was<br />

difficult to maintain this interval due to steep slopes<br />

or thick vegetation, we were able only to<br />

approximate it). Previously identified sites were<br />

collected more intensively, using either a systematic<br />

grid <strong>of</strong> 10 m 2 collection areas or nonsystematic<br />

collection <strong>of</strong> any artifacts observed on the site. In<br />

1997 and 1998 some areas with previously reported


sites were surveyed less systematically, but we<br />

attempted for the sake <strong>of</strong> comparability to use<br />

systematic methods for all areas in 1999.<br />

The 1999 survey was conducted from 29<br />

August to 17 September. This included 2 days <strong>of</strong><br />

fieldwalking with two crews, 5 days <strong>of</strong> fieldwalking<br />

with one crew, 0.5 days <strong>of</strong> fieldwalking with a<br />

double-size crew, and 3.5 days <strong>of</strong> intensive surface<br />

collection with a large crew at Mazza. In all this<br />

represents 73 person-days <strong>of</strong> work, <strong>of</strong> which 45.5<br />

were used for fieldwalking and 27.5 for intensive<br />

collection at Mazza.<br />

The fieldwalking survey in 1999 covered a<br />

total <strong>of</strong> 102.0 ha, <strong>of</strong> which 5.8 ha had been<br />

surveyed in previous years and 96.2 ha were new.<br />

In all, excluding repeat visits to the same area, we<br />

have surveyed a total <strong>of</strong> 274.5 ha in the three years<br />

<strong>of</strong> the project so far. We defined 53 new survey<br />

areas in 1999, numbered 102 through 154. Some <strong>of</strong><br />

these (Areas 102, 103, 130, 131, 133, and 134) may<br />

overlap to some degree with areas from previous<br />

years. All new areas were surveyed using<br />

systematic transect walking except Areas 150<br />

through 154, which were areas at Mazza surveyed<br />

using a systematic grid <strong>of</strong> collection areas instead<br />

(see below).<br />

The survey work <strong>of</strong> the past three years<br />

has been concentrated in several locations. The two<br />

largest clusters are around Umbro (Figure 1, Figure<br />

2) and in the lower part <strong>of</strong> the San Pasquale valley.<br />

In each <strong>of</strong> these places numerous survey areas make<br />

up almost a square kilometer <strong>of</strong> contiguous<br />

coverage. Most <strong>of</strong> this was done in 1997 and 1998,<br />

but a few areas were added at Umbro in 1999,<br />

especially to the northeast. A third cluster <strong>of</strong> survey<br />

areas is located in the middle part <strong>of</strong> the San<br />

Pasquale valley, but in this case it consists <strong>of</strong><br />

several disconnected fragments, due to difficulties<br />

<strong>of</strong> access. A few <strong>of</strong> these areas were done in 1997<br />

and 1998, but most were done in 1999. Several<br />

smaller clusters have been selected to represent<br />

particular types <strong>of</strong> location: Mazza (Figure 3) and<br />

Capo Crisafi for coastal hills (1997, plus the<br />

controlled collections in 1999), M. Rotonda and M.<br />

Vunemo for high inland hills (both newly done in<br />

1999), and M. Silipone to compare the valley <strong>of</strong> the<br />

Torrente Sideroni with the San Pasquale valley<br />

(mostly done in 1998, plus two small areas in<br />

1999). Several small isolated patches occur as well,<br />

mostly to investigate known sites or to survey small<br />

patches <strong>of</strong> accessible land near the modern town <strong>of</strong><br />

<strong>Bova</strong> <strong>Marina</strong>; a few <strong>of</strong> these have been done each<br />

year.<br />

10<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

2.1.3. Ceramics and chronology<br />

One <strong>of</strong> the biggest problems for survey in<br />

the past was the lack <strong>of</strong> a well-defined local<br />

ceramic sequence which would be used to date sites<br />

found on survey. During 1999, considerable<br />

progress was made at identification <strong>of</strong> the pottery<br />

from the survey. We were able to work several days<br />

in the Museo Archeologico Nazionale in Reggio di<br />

Calabria before the field season, restudying some <strong>of</strong><br />

our 1997 and 1998 collections and also comparing<br />

them to excavated materials from the Roman site at<br />

San Pasquale (Deri). We are grateful to Dssa.<br />

Emilia Andronico and the staff <strong>of</strong> the museum for<br />

their assistance with this work.<br />

Based on this research, a new provisional<br />

ceramic classification was set up before the field<br />

season (and modified slightly during the field<br />

season), which was used for classification <strong>of</strong> the<br />

1999 survey finds. This system will require<br />

additional modification as our knowledge <strong>of</strong> the<br />

regional pottery improves, but it has improved our<br />

ability to recognize chronological information in<br />

our surface collections, by enabling us to associate<br />

the better-known wares with changes in the local<br />

productions.<br />

Of the remaining problems in this ceramic<br />

chronology, the most important is the prehistoric<br />

period. Prehistoric pottery can usually be<br />

distinguished from that <strong>of</strong> later periods, but that is<br />

<strong>of</strong>ten the limit <strong>of</strong> our resolution so far. The standard<br />

chronological types for prehistoric pottery in<br />

southern Italy, based on whole vessels from burials,<br />

make use <strong>of</strong> both form and decoration. For some<br />

periods, such as the Neolithic, surface decoration is<br />

<strong>of</strong>ten useful, but after the Neolithic, most vessels<br />

were probably undecorated and vessel form is the<br />

most distinctive criterion. Unfortunately, the small<br />

eroded fragments in our surface collections have<br />

little evidence for either, but especially for vessel<br />

form. Traces <strong>of</strong> impressed decoration sometimes<br />

survive, making the Neolithic period more visible<br />

than other portions <strong>of</strong> prehistory, while post-<br />

Neolithic periods are far less easy to identify. It is<br />

likely that some very broad divisions should be<br />

possible on the basis <strong>of</strong> fabric and surface<br />

treatment, although these are not likely to be as<br />

precise as the existing categories. Of our prehistoric<br />

sites, some have been dated by chance survival <strong>of</strong> a<br />

diagnostic element or two, some have a suggested<br />

date based on a subjective assessment <strong>of</strong> similarity<br />

to the excavated assemblage from Umbro, and<br />

some remain undated.<br />

The chronology <strong>of</strong> the Greek period is<br />

founded on the black-gloss finewares which,<br />

although mostly <strong>of</strong> regional manufacture, reflect<br />

stylistic trends common throughout the Greek<br />

world. These are associated with reddish brown


sandy coarseware, beige plainwares, large and<br />

coarse storage jars, and sandy beige transport<br />

amphoras <strong>of</strong> Greek-derived form. Ro<strong>of</strong> tile appears<br />

for the first time in the Greek period as well,<br />

predominantly in a sandy, light-colored fabric and<br />

taking the form <strong>of</strong> flat tile (tegula) with wide<br />

flanges and cover tile (imbrex) with a flat top and<br />

angular sides.<br />

The assemblages <strong>of</strong> the<br />

Hellenistic/Republican and early Imperial periods<br />

remain ill-defined, because very few imports <strong>of</strong><br />

well-dated types have been found to confirm dates<br />

in this range. The assemblages <strong>of</strong> local<br />

commonwares seem to show a gradual changing<br />

mixture, however, <strong>of</strong> the sandy brown cooking ware<br />

and fine plainware <strong>of</strong> the Greek period and the<br />

gritty light-colored wares <strong>of</strong> the later Roman<br />

period. An orange variant <strong>of</strong> the Greek cooking<br />

ware is likely to belong to this period.<br />

The late Roman period is well defined by<br />

the presence <strong>of</strong> imports, including African Red Slip<br />

ware, plainwares, and amphoras from North Africa,<br />

as well as small quantities <strong>of</strong> Late Roman C ware<br />

and Late Roman amphoras from the eastern<br />

Mediterranean. The types and quantities found in<br />

<strong>Bova</strong> <strong>Marina</strong> suggest a peak <strong>of</strong> imports between the<br />

late 2nd or early 3rd and late 5th centuries AD. The<br />

associated local wares include light-colored<br />

coarseware and amphoras tempered with wellsorted<br />

grit (around 1 mm in size) and light-colored<br />

fine plainware. Roman ro<strong>of</strong> tile, compared to<br />

Greek, tends to be in grittier fabrics and to have<br />

slightly different forms, with thinner flanges on the<br />

tegulae and a curved shape to the imbrices.<br />

Probably at the later end <strong>of</strong> the sequence <strong>of</strong> Roman<br />

coarsewares is a brown fabric tempered with<br />

abundant grit, mica, and sometimes chunks <strong>of</strong><br />

schist. The precision <strong>of</strong> dates within the Roman<br />

period is limited so far, however, because the sites<br />

appear in most cases to have been occupied for<br />

quite long periods, several centuries in the case <strong>of</strong><br />

Sites 9 and 22.<br />

There is a large, ill-defined gap between<br />

the end <strong>of</strong> the Roman period and the recent<br />

assemblage associated with farmhouses <strong>of</strong> the 19th<br />

and 20th centuries. It is not clear how much <strong>of</strong> this<br />

results from an actual scarcity <strong>of</strong> settlement and<br />

how much from poor recognition <strong>of</strong> the pottery <strong>of</strong><br />

those periods. Judging from the few available<br />

reports, from the 9th to 11th centuries, ceramic<br />

assemblages in Calabria seem to consist <strong>of</strong> coarse<br />

cooking wares, semi-fine or finely sandy wares with<br />

red painted decoration, and Italian-Byzantine<br />

amphoras. All <strong>of</strong> these are continuations from Late<br />

Antiquity, although the details may vary. In<br />

addition, occasional lead-glazed pottery begins by<br />

the 9th century, either with a thick green glaze or<br />

11<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

with a sparse glaze overall but large patches and<br />

streaks <strong>of</strong> thick glaze. In the 11th century the first<br />

tin-glazed wares with green and brown decoration<br />

appear. Very little has been published on late<br />

medieval pottery in Calabria, but based on other<br />

parts <strong>of</strong> southern Italy one would expect tin-glazed<br />

protomaiolica decorated in various colors, sgraffito<br />

wares, lead-glazed cooking wares, and unglazed<br />

semi-fine wares with red painted decoration<br />

(narrower, finer painting than on the early medieval<br />

version). Until a better sense <strong>of</strong> the local<br />

productions can be obtained, it may be possible for<br />

early medieval pottery to be mistaken for late<br />

ancient, and for late medieval pottery to be<br />

mistaken for early modern.<br />

There is a clearly modern assemblage,<br />

<strong>of</strong>ten associated with abandoned farmhouses,<br />

including large beige water jars, red casseroles with<br />

a very thin green glaze, beige plates, bowls, and<br />

jars with a thin greenish or brownish glaze,<br />

earthenwares with decorated opaque glazes, and<br />

hard, well-fired, usually dark red curved ro<strong>of</strong> tiles.<br />

This assemblage probably dates to the late 19th and<br />

20th centuries. Some <strong>of</strong> our collections resemble<br />

this assemblage in some respects but also differ<br />

substantially. These collections may be associated<br />

with visible ruins, as at Site 10 (Torre Crisafi), but<br />

<strong>of</strong>ten are not. Characteristics include red casseroles<br />

decorated with white slip under a yellow glaze, jars<br />

in a hard, finely granular reddish fabric, and curved<br />

ro<strong>of</strong> tile in a s<strong>of</strong>t brown fabric containing large<br />

chunks <strong>of</strong> gneiss and schist. In some cases, such as<br />

Torre Crisafi, there is good reason to assign this<br />

material to the early modern period, but in most<br />

cases there is no evidence for an absolute date, and<br />

some <strong>of</strong> it may extend back to the late Middle<br />

Ages. Comparison with better-dated assemblages<br />

from elsewhere will therefore be important for<br />

establishing the chronological ranges <strong>of</strong> these types.<br />

2.1.4. Results to date: sites found<br />

We assigned eight new site numbers in<br />

1999, bringing the total to 37 (Figure 1). We<br />

recognize, however, that not all <strong>of</strong> these site<br />

numbers represent comparable entities. They range<br />

from a few scattered artifacts in a field all the way<br />

to a large and dense concentration <strong>of</strong> finds such as<br />

Mazza (Site 12). The issue is further complicated<br />

by the different levels <strong>of</strong> material culture typical <strong>of</strong><br />

different periods: a thin scatter <strong>of</strong> artifacts which<br />

would define a full-fledged prehistoric site might be<br />

less than the general <strong>of</strong>f-site “background” scatter<br />

<strong>of</strong> artifacts in areas <strong>of</strong> intensive Roman land use.<br />

As a first approximation toward a more<br />

useful classification, we have divided them into two<br />

categories: "sites", meaning locations where the


artifact concentration is sufficiently obvious that<br />

one can define its extent, and "scatters", meaning<br />

areas or locations where artifacts are more<br />

abundant than usual, but not abundant enough to<br />

form a definite concentration.<br />

The list here reviews all 37 locations to<br />

which we have given site numbers so far. Note that,<br />

for the sites defined in 1997 and 1998, it<br />

incorporates updated information where available<br />

from revisiting the location or restudying the<br />

collections. Some sites have been redated or<br />

reinterpreted, and hence this information<br />

supersedes earlier descriptions.<br />

Site 1 (Canturatta A, Area 2). Site, ca. 1<br />

ha? Neolithic. This site, reported to us by S.<br />

Stranges and L. Saccà , is located on the eastern<br />

edge <strong>of</strong> the S. Pasquale valley, on the slopes just<br />

below a steep ridge. It is in a field with some vines<br />

and almond trees, and evidence for more extensive<br />

vine cultivation in the past. The 1997 collections<br />

from this area produced one Impressed Ware or<br />

Stentinello sherd, and Stranges reports having<br />

found several Stentinello sherds there. Test<br />

excavations in 1999, described in detail elsewhere<br />

in this report, failed to demonstrate the existence <strong>of</strong><br />

significant subsurface archaeological deposits.<br />

Site 2 (Canturatta B, Areas 2, 50). Site, ca.<br />

2 ha? Roman. This site is located on the mild slopes<br />

on the eastern edge <strong>of</strong> the lower S. Pasquale valley,<br />

just below a steep ridge. Abundant fragments <strong>of</strong><br />

Roman commonwares and tile were found,<br />

especially in the northeastern corner <strong>of</strong> Area 50.<br />

The collections need to be reexamined, but no<br />

clearly datable diagnostics have been noted so far.<br />

Site 3 (Pisciotta A, Area 5). Site, ca. 0.5<br />

ha. Bronze Age. This site consists <strong>of</strong> a scatter <strong>of</strong><br />

impasto sherds found on a steep rocky hillside at<br />

the eastern edge <strong>of</strong> the S. Pasquale valley. The<br />

artifacts appear to be associated with a particular<br />

level near the base <strong>of</strong> the hillside, are highly<br />

localized, and include fairly large pieces <strong>of</strong> vessels.<br />

This implies that the site is located on the slope<br />

itself rather than consisting <strong>of</strong> slopewash from<br />

above. The pottery includes pieces with carination,<br />

high raised strap handles, or horizontal lug handles,<br />

suggesting a Bronze Age date.<br />

Site 4 (Pisciotta B, Area 6). Scatter.<br />

Roman? This area is located near the southern end<br />

<strong>of</strong> the top <strong>of</strong> the ridge overlooking the eastern side<br />

<strong>of</strong> the S. Pasquale valley. A scatter <strong>of</strong> ancient tile<br />

and pottery fragments, possibly Roman, was found<br />

in this area, in contrast to adjoining areas, but the<br />

quantity <strong>of</strong> material found is small enough that it<br />

may be background scatter associated with a site<br />

elsewhere.<br />

12<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

Site 5 (Pisciotta C, Area 7). Site, 0.4 ha.<br />

Bronze Age? This site, reported to us by S.<br />

Stranges and L. Saccà, is located on a rocky hill<br />

along the ridge to the east <strong>of</strong> the S. Pasquale valley.<br />

A scatter <strong>of</strong> prehistoric pottery was found on the<br />

peak <strong>of</strong> the hill and just below it on the western<br />

side. Most <strong>of</strong> this pottery was nondiagnostic, but<br />

one rim would be compatible with a Bronze Age or<br />

Iron Age date, and previous collections described<br />

by Saccà would support a Bronze Age date. If so,<br />

like Site 3, it may be worth further investigation to<br />

see how similar it is to recently discovered BA<br />

deposits at the base <strong>of</strong> a rocky bank at Umbro (see<br />

below).<br />

Site 6 (Deri A, Areas 9, 10, 83). Site, ca. 4<br />

ha. Roman. This site is just above the east bank <strong>of</strong><br />

the Fiumara di S. Pasquale near the sea. A rescue<br />

excavation in advance <strong>of</strong> partially completed<br />

highway construction several years ago found a<br />

Late Roman synagogue as well as other structures<br />

and some burials (Costamagna 1991). The principal<br />

data from our survey consist <strong>of</strong> a series <strong>of</strong> 15<br />

irregularly placed 10 m 2 collection units in the field<br />

just north <strong>of</strong> the intended highway bridge (Area 10)<br />

as well as collections from transect-walking in<br />

adjoining areas. In addition, a few diagnostic pieces<br />

were selected from the disturbed area around the<br />

construction. There is at least a thin scatter <strong>of</strong><br />

artifactual material throughout the entire field, but a<br />

dense concentration only in the southwestern half<br />

<strong>of</strong> Area 10. The pottery appears to cover a time<br />

range at least from the 1st century BC to the 5th<br />

century AD. Dated sherds include two rims <strong>of</strong> a<br />

form resembling Campana A Lamboglia 36 with a<br />

grayish brown slip (2nd to 1st century BC), a<br />

fragment <strong>of</strong> a thin-walled cup decorated with<br />

barbotine dots (late 1st century BC to early 2nd<br />

century AD), Eastern Sigillata B Hayes 60 (80-150<br />

AD), African Red Slip (ARS) Hayes 8A (75/90-<br />

180/200 AD), ARS cookware Hayes 23B (150-220<br />

AD), African amphora Keay 27B (300-450 AD),<br />

ARS Hayes 50B (350-400+ AD), ARS Hayes 67<br />

(350-450 AD), and ARS Hayes 84 small (440-500<br />

AD). Additional finds include Campanian (black<br />

sand) amphora, Late Roman Amphora 2, and a<br />

marble mosaic tessera.<br />

Site 7 (Deri B, Areas 9, 84). Site, ca. 1 ha.<br />

Prehistoric. This site, which overlaps the northern<br />

part <strong>of</strong> the Roman site at Deri, is a sparse scatter <strong>of</strong><br />

prehistoric artifacts in a cultivated field in the<br />

bottom <strong>of</strong> the S. Pasquale valley. The finds include<br />

a small amount <strong>of</strong> prehistoric impasto, and one<br />

fragment <strong>of</strong> obsidian. There was also a piece <strong>of</strong><br />

worked flint nearby in Area 10. The only diagnostic<br />

artifact was a horizontal lug handle. This site is<br />

noteworthy for being a prehistoric site in a valley<br />

bottom location, suggesting that landscape<br />

alteration in the valleys has not been so total as to


preclude all possibility <strong>of</strong> finding sites. It may be<br />

the one erroneously called “Torre Varata” in Tinè<br />

(1992).<br />

Site 8 (Pisciotta D, Area 13). Scatter.<br />

Historic. This is a sparse scatter <strong>of</strong> pottery found in<br />

a small area about 30 meters in diameter on a<br />

recently reforested slope. The artifacts are not<br />

prehistoric, but have yet to be reexamined to<br />

determine their date.<br />

Site 9 (Umbro A, Area 16; Figure 2)).<br />

Site, < 0.1 ha. Neolithic, Copper Age?, Bronze<br />

Age, Roman. This site has been the principal focus<br />

<strong>of</strong> the excavations in 1998 and 1999 and is<br />

discussed in detail in elsewhere in this report. To<br />

summarize briefly, this site, previously explored by<br />

S. Stranges and briefly examined by Tinè (1992),<br />

is located on a plateau near the border between<br />

<strong>Bova</strong> <strong>Marina</strong> and <strong>Bova</strong> Superiore, where the<br />

bedrock <strong>of</strong> calcareous sandstone projects out to<br />

form steep cliffs. At the foot <strong>of</strong> one <strong>of</strong> the cliffs was<br />

found a dense scatter <strong>of</strong> Neolithic pottery and<br />

obsidian; this area was excavated in both 1998 and<br />

1999. Two areas on top <strong>of</strong> this same cliff where<br />

undated prehistoric pottery had been found on the<br />

surface were also excavated, one in 1998 yielding a<br />

poorly dated assemblage, possibly post-Neolithic,<br />

and the other in 1999 yielding an assemblage<br />

including parts <strong>of</strong> some Early Bronze Age whole<br />

vessels. Abundant small pieces <strong>of</strong> human bone were<br />

found partway up the cliff; although associated with<br />

Neolithic pottery and obsidian, these were<br />

subsequently dated by radiocarbon to the late<br />

Roman period. The presence <strong>of</strong> Roman burials at<br />

this location is somewhat enigmatic, because the<br />

nearest site found as yet which may possibly have<br />

Roman occupation is Site 33, about half a kilometer<br />

away. However, Area 130, located about 250<br />

meters to the southwest, yielded one fragment <strong>of</strong><br />

Italian sigillata, so there is some other evidence for<br />

Roman activity in the area.<br />

Site 10 (Torre Crisafi, Area 26). Site, ca.<br />

0.5 ha. Modern. This site occupies the peak <strong>of</strong> the<br />

promontory above Capo S. Giovanni. The central<br />

part <strong>of</strong> the peak is now occupied by the shrine <strong>of</strong><br />

the Madonna del Mare. However, around the edges<br />

<strong>of</strong> the peak and around the ruined coastal<br />

watchtower, abundant pottery and tile fragments<br />

were collected. Although this site was originally<br />

considered both Classical and medieval in date,<br />

part <strong>of</strong> the collection was reexamined in 1999 and<br />

the site has been redated. It appears to be modern,<br />

but probably earlier than the assemblage associated<br />

with recent abandoned farms in <strong>Bova</strong> <strong>Marina</strong>. It<br />

might be useful to compare this assemblage with<br />

late medieval pottery as well. It is worth noting that<br />

the tower at this point dates to the 16 th -17 th<br />

centuries.<br />

13<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

Site 11 (Cimitero, Areas 28, 29). Site?<br />

Historic. A large amount <strong>of</strong> pottery was found on<br />

the slope west <strong>of</strong> the modern Cimitero S. Pietro,<br />

overlooking the modern town <strong>of</strong> <strong>Bova</strong> <strong>Marina</strong>. A<br />

small amount was also found at the top <strong>of</strong> the hill.<br />

The artifacts are not prehistoric, but have yet to be<br />

reexamined to determine their date.<br />

Site 12 (Mazza, Areas 30-37, 102, 103,<br />

150-153; Figure 3)). Site, ca. 8 ha. This site,<br />

reported to us by S. Stranges and L. Saccà and the<br />

location several years ago <strong>of</strong> a small test excavation<br />

by L. Costamagna for the Soprintendenza<br />

Archeologica della Calabria, was partially surveyed<br />

in 1997. At that time the presence <strong>of</strong> a substantial<br />

Greek site, presumably a village, was confirmed.<br />

Restudy <strong>of</strong> the collections in 1999 showed the<br />

presence <strong>of</strong> prehistoric and Late Roman pottery in<br />

addition to a wide range <strong>of</strong> Greek pottery, from<br />

early Archaic to Hellenistic. Greek pottery included<br />

the base <strong>of</strong> an early Archaic cup with bichrome<br />

decoration, numerous cup and krater pieces <strong>of</strong> the<br />

6th to 5th centuries BC, Italiote transport amphoras<br />

<strong>of</strong> the 5th to 4th centuries BC, and a possible<br />

Corinthian “frying pan” <strong>of</strong> the Hellenistic period.<br />

Late Roman material identified in the 1997<br />

collections includes ARS Hayes 91C/D (6th to 7th<br />

century AD), ARS Hayes 105 (late 6th to 7th<br />

century AD), and Late Roman C Hayes 3C/D (late<br />

5th century AD). In 1999, we did a small amount <strong>of</strong><br />

new fieldwalking (Areas 102 and 103) and<br />

conducted intensive surface collections on a regular<br />

grid over all the parts <strong>of</strong> the site to which we could<br />

obtain access. These intensive collections are<br />

reported in detail below. It should also be noted<br />

that the site probably continues to the east, and<br />

possibly to the north as well, but that access to that<br />

area was blocked by fences in both 1997 and 1999.<br />

Site 13 (Pisciotta E, Area 47). Site, ca. 0.5<br />

ha. Roman. This site was reported to us as a<br />

possible Roman kiln site by S. Stranges and L.<br />

Saccà . It is located in a small ravine cut into a<br />

slope on the eastern edge <strong>of</strong> the S. Pasquale valley.<br />

On the slope north <strong>of</strong> the ravine, abundant tile and<br />

commonware sherds were found. A drystone wall<br />

foundation, possibly <strong>of</strong> a small rectangular<br />

building, was observed in the eroding hillside. Two<br />

rounded depressions in the bottom <strong>of</strong> the ravine<br />

may be the features suggested to have been kilns.<br />

Full reexamination <strong>of</strong> the collection remains to be<br />

done, but a brief inspection showed a small but<br />

diverse assemblage <strong>of</strong> Roman (and possibly also<br />

Greek or medieval) commonwares, including a<br />

fragment <strong>of</strong> African amphora. Surface collections<br />

contain no evidence <strong>of</strong> pottery production such as<br />

misfired wasters or a predominance <strong>of</strong> one or two<br />

fabrics. Thus, it may be best interpreted as a typical<br />

small Roman site, probably a farm. The round


features <strong>of</strong> unknown date and function may or may<br />

not be associated.<br />

Site 14 (Panaghia A, Areas 51, 55). Site,<br />

ca. 0.5 ha? Roman. This site, reported to us by S.<br />

Stranges and L. Saccà , is located on the western<br />

side <strong>of</strong> the middle part <strong>of</strong> the San Pasquale valley.<br />

Collections from Area 51, done in 1997, include a<br />

variety <strong>of</strong> Roman commonwares and tile, probably<br />

predominantly early imperial, and a fragment <strong>of</strong> a<br />

fineware with a bright orange slip (possibly either<br />

ARS chiara A or Eastern Sigillata B). When the<br />

site was revisited in 1998, a trench had been dug in<br />

the site, apparently for construction, which had cut<br />

through a Roman structure with tile ro<strong>of</strong>, brick<br />

walls, and opus signinum (cocciopesto) floor before<br />

being halted. In the disturbed ground around this<br />

trench we collected more Roman artifacts,<br />

including the base <strong>of</strong> an ARS bowl (chiara D<br />

fabric) with feather-rouletting on the interior, dating<br />

to the 5th or 6th century AD. We made a series <strong>of</strong><br />

controlled collections in the field extending north<br />

<strong>of</strong> this trench (Area 55). For the most part these<br />

collections produced little or no Roman material,<br />

except for a low-density scatter about 100 meters<br />

away from the structure. Thus, the site is apparently<br />

fairly small, probably a single farm. The area to the<br />

east is inaccessible, being fenced-<strong>of</strong>f orange groves,<br />

but it may be worthwhile to investigate how far the<br />

artifact scatter extends to the southwest.<br />

Site 15 (Agrillei, Area 52). Site, < 1 ha.<br />

Prehistoric. This site is located on a small<br />

promontory at the southern end <strong>of</strong> the Agrillei ridge<br />

near the border between <strong>Bova</strong> <strong>Marina</strong> and Palizzi<br />

<strong>Marina</strong>, overlooking the sea. The presence <strong>of</strong> a<br />

prehistoric site here was reported to us by L. Saccà.<br />

Collections in 1997 and 1998 produced a<br />

significant quantity <strong>of</strong> prehistoric impasto sherds,<br />

mostly from the southern face <strong>of</strong> the hill, just below<br />

the crest. None <strong>of</strong> the fragments appear to be useful<br />

chronological indicators. The presence <strong>of</strong> a flat area<br />

at the summit and buried terrace walls eroding out<br />

<strong>of</strong> the slope suggests modern earth-moving, so the<br />

finds may not be in their original context. Several<br />

probably Greek or Roman sherds indicate the<br />

possibility <strong>of</strong> a Greek or Roman component as well,<br />

but the quantity is not enough to rule out their being<br />

background scatter.<br />

Site 16 (Umbro B, Area 58). Site, ca. 0.4<br />

ha. Prehistoric. This site is on a hill about 150<br />

meters north <strong>of</strong> Site 9 at Umbro. The main<br />

concentration <strong>of</strong> artifacts is on a small saddle <strong>of</strong><br />

land along the top <strong>of</strong> the hill and along the upper<br />

northern face <strong>of</strong> the hill. Collections in 1998<br />

produced a large quantity <strong>of</strong> prehistoric impasto;<br />

the assemblage differs somewhat from the Neolithic<br />

pottery at Site 9, particularly in the presence <strong>of</strong> a<br />

slip on many fragments, but no obvious<br />

14<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

chronological diagnostics were found. It is likely<br />

that the assemblage is post-Neolithic, but<br />

excavation may be needed to determine a more<br />

specific date.<br />

Site 17 (Limaca, Area 66). Scatter.<br />

Prehistoric. On the northern slope <strong>of</strong> a rounded hill<br />

east <strong>of</strong> Umbro we collected one impasto sherd and<br />

one piece <strong>of</strong> worked flint. The quantity is too small<br />

to demonstrate the existence <strong>of</strong> any concentration<br />

that could be called a site, but ground visibility was<br />

poor, so it may be that additional items were<br />

missed.<br />

Site 18 (Umbro C, Areas 24, 68, 130).<br />

Site, 0.2 ha. Greek. The main concentration <strong>of</strong> finds<br />

is in Area 24, which is the sloping top <strong>of</strong> a small<br />

rocky outcrop. A less dense scatter <strong>of</strong> artifacts<br />

occurs also in the adjoining areas. The finds include<br />

black-gloss pottery, probably <strong>of</strong> the late 6th to early<br />

5th century BC, as well as Greek commonwares<br />

and tile. The small but fairly dense sherd scatter<br />

suggests something like a small single farm site. A<br />

few Roman sherds have been found in nearby areas<br />

as well; it is not clear whether these relate to reuse<br />

<strong>of</strong> this location or to some other, as yet unidentified<br />

site.<br />

Site 19 (Penitenzeria, Area 72). Scatter.<br />

Prehistoric?, Greek. This area is a large, gently<br />

sloping field on the large plateau extending west<br />

and south from the rock outcrops at Umbro. A<br />

small scatter <strong>of</strong> artifacts was collected from the<br />

southern edge <strong>of</strong> Area 72. This included four<br />

probable fragments <strong>of</strong> prehistoric impasto, all<br />

nondiagnostic, as well as a fragment <strong>of</strong> Greek<br />

black-gloss and some Greek or Roman<br />

commonwares. This location should be revisited to<br />

assess whether there is a significant artifact<br />

concentration, and if so to determine its extent and<br />

obtain more evidence <strong>of</strong> its date.<br />

Site 20 (Buccisa A, Area 76). Scatter.<br />

Greek/Roman? This area is an olive grove on the<br />

south face <strong>of</strong> M. Buccisa. At the southern edge, a<br />

dense concentration <strong>of</strong> Greek or Roman tile was<br />

found. No definitely ancient pottery was found in<br />

association with this tile. Thus, it is uncertain<br />

whether this tile scatter is the result <strong>of</strong> an ancient<br />

structure in this location or <strong>of</strong> reuse for building<br />

modern terrace walls. Even in the latter case,<br />

however, it is likely that the site the tile came from<br />

should be nearby.<br />

Site 21 (Buccisa B, Area 79). Scatter.<br />

Prehistoric. In a small, level area at the foot <strong>of</strong> the<br />

southwest end <strong>of</strong> M. Buccisa, a small scatter <strong>of</strong><br />

pottery was found in a cultivated field. Most was<br />

modern, but three fragments <strong>of</strong> prehistoric impasto,<br />

all nondiagnostic, were also present. This is unusual<br />

for such a small area, and although this small


quantity could be background scatter, it is also<br />

possible that there is a prehistoric site in this<br />

location or nearby.<br />

Site 22 (Sideroni/Amigdala, Areas 88, 89).<br />

Site, > 1.5 ha. Prehistoric?, Roman, Medieval. The<br />

site is located on the eastern bank <strong>of</strong> the Torrente<br />

Sideroni, amidst the built-up area <strong>of</strong> modern <strong>Bova</strong><br />

<strong>Marina</strong>. Much <strong>of</strong> Area 88 is directly under a<br />

highway overpass. Collections were made from<br />

vacant lots in Area 89 and the northern part <strong>of</strong> Area<br />

88 and from disturbed areas under the highway<br />

overpass and near recent construction in Area 88.<br />

These disturbed areas produced very abundant<br />

finds in good condition; the vacant lots, despite<br />

poor visibility due to grassy ground cover,<br />

produced moderately abundant finds as well. The<br />

total size <strong>of</strong> the site is impossible to estimate,<br />

because only a few places are not currently built<br />

over. No clear evidence was found regarding<br />

whether the site should be considered as a village<br />

or as a villa. The site was under excavation by the<br />

Soprintendenza Archeologica della Calabria in<br />

1999; it is hoped that their investigation will enable<br />

better interpretation <strong>of</strong> the nature <strong>of</strong> the site's<br />

occupation. The earliest well-dated sherds were a<br />

fragment <strong>of</strong> a thin-walled cup, perhaps <strong>of</strong> the 1st<br />

century AD, and the rim <strong>of</strong> a bowl in a gray fabric<br />

covered with a dark gray slip, possibly a Sardinian<br />

or Sicilian imitation <strong>of</strong> Campanian ware <strong>of</strong> the 2nd<br />

or 1st century BC. The principal concentration <strong>of</strong><br />

finds ranges from the 2nd to the 5th century AD.<br />

ARS forms include Hayes 23, Hayes 32/58, Hayes<br />

52B, Hayes 196, Hayes 59, Hayes 61, Lamboglia<br />

52B, and a small fragment <strong>of</strong> a decorated lamp.<br />

Several pieces <strong>of</strong> African amphora were found,<br />

including an Africana II spike and a Keay 62a rim.<br />

Identified eastern Mediterranean imports include<br />

Late Roman Amphoras 2 and 3. Some red-painted<br />

commonware sherds <strong>of</strong> late ancient or early<br />

medieval date were found. Early medieval<br />

occupation was demonstrated by the presence <strong>of</strong> a<br />

lamp <strong>of</strong> the so-called lucerna a ciabatta type, dated<br />

approximately to the 8th century AD (Garcea and<br />

Williams 1987; Ceci 1992). Four possibly<br />

prehistoric impasto sherds were found in the<br />

northern part <strong>of</strong> the site (Area 89), but they were<br />

nondiagnostic.<br />

Site 23 (Carusena, Area 91). Scatter. Two<br />

prehistoric sherds were found in this small area in<br />

1998. Given the size and location <strong>of</strong> the area this<br />

was unexpected, so a site number was given to this<br />

scatter. In 1999 part <strong>of</strong> the ridge just above Area 91<br />

was walked (Area 148), yielding no prehistoric<br />

artifacts. Although part could not be done at the<br />

time, it is likely that there is no concentrated<br />

prehistoric site in the vicinity, and that the minor<br />

scatter in Area 91 is as much as will be found.<br />

Although one <strong>of</strong> the sherds found in 1998 was a<br />

15<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

rim, it did not appear to be chronologically<br />

diagnostic.<br />

Site 24 (Zaccaria, Area 96). Scatter.<br />

Roman. This area is on the middle part <strong>of</strong> the slope<br />

<strong>of</strong> M. Silipone, overlooking the valley <strong>of</strong> the<br />

Torrente Sideroni. This part <strong>of</strong> the slope yielded<br />

several fragments <strong>of</strong> ancient pottery and tile,<br />

probably Roman. There was no clear concentration<br />

anywhere, but the quantity is much greater than in<br />

the surrounding areas. The quantity is not enough to<br />

rule out being background scatter related to an as<br />

yet unlocated site, but because the area is located<br />

on a s<strong>of</strong>t clay slope, it is also possible that a small<br />

site here has been largely obscured by<br />

geomorphological processes.<br />

Site 25 (Sant'Aniceto, Area 98). Site, ca.<br />

0.6 ha. Prehistoric (Final Bronze Age or Iron<br />

Age?), Roman/Medieval?, Modern. This site is<br />

located at the top <strong>of</strong> a very steep hill overlooking<br />

the valley <strong>of</strong> the Torrente Vena. The presence <strong>of</strong> a<br />

medieval church on this hill was reported to us by<br />

S. Stranges and L. Saccà. A difficult and timeconsuming<br />

climb is required to get to the top <strong>of</strong> the<br />

hill. The most obvious feature at the top is a small<br />

ruined church built <strong>of</strong> mortared stone masonry,<br />

approximately 9-10 meters square with an apse on<br />

the eastern side. Some other traces <strong>of</strong> wall<br />

foundations are visible several meters to the north.<br />

The predominant finds include red straw-tempered<br />

ro<strong>of</strong> tile, red earthenware with a thick green glaze,<br />

light-colored plainware with reddish sherd temper,<br />

and light-colored to orange-brown gritty<br />

coarseware. The last <strong>of</strong> these resembles Roman<br />

coarsewares, but no datable finewares or imports<br />

were found. The others are unusual in our<br />

collections and appear most likely to be early<br />

modern, or possibly late medieval, although no<br />

recognizably medieval types were found. On the<br />

present evidence the church is most likely to be<br />

modern, although an earlier date cannot be ruled<br />

out. In the eroding slope just west <strong>of</strong> the church a<br />

somewhat different pottery assemblage was found,<br />

consisting mostly <strong>of</strong> impasto, probably <strong>of</strong><br />

protohistoric date. Three decorated pieces <strong>of</strong> coarse<br />

pottery were found, one with a wide band <strong>of</strong> dark<br />

brown paint and two with finger-impressed applied<br />

cordons; these could be <strong>of</strong> either protohistoric (Iron<br />

Age?) or medieval date.<br />

Site 26 (Climarda, Area 99). Site,<br />

unknown extent. Greek?, Roman. This area consists<br />

<strong>of</strong> the fields at the foot <strong>of</strong> the south end <strong>of</strong> the hill<br />

on which Site 25 is located, as well as the banks <strong>of</strong><br />

the Torrente Vena running through these fields.<br />

Most <strong>of</strong> the finds in this area were modern,<br />

probably related to an abandoned farmhouse at the<br />

foot <strong>of</strong> the hill. In the eroding river banks, however,<br />

and archaeological stratum could be seen about 1 to


1.5 meters below the present ground surface. From<br />

the banks and the stream bed, and also in small<br />

quantities from the fields, we recovered numerous<br />

fragments <strong>of</strong> ancient pottery and tile, including<br />

black-gloss (Greek or Republican), ARS form<br />

Hayes 96 (early 6th century AD), and the spike <strong>of</strong> a<br />

Late Roman African amphora. It is not clear from<br />

the small collection made in 1998 whether the<br />

apparent hiatus between the Greek/Republican<br />

occupation and the Late Roman is real or not. Due<br />

to the depth <strong>of</strong> the overburden it is difficult to<br />

determine the extent <strong>of</strong> the site from the surface<br />

evidence, but the depth <strong>of</strong> the deposit also suggests<br />

that this may be a rare site with relatively<br />

undisturbed archaeological contexts. Thus, it may<br />

merit further archaeological testing, at least to the<br />

extent <strong>of</strong> cleaning the bank pr<strong>of</strong>iles.<br />

Site 27 (Vadicamo, Area 100). Site, < 0.01<br />

ha. Roman/Medieval. Area 100 is a long, badly<br />

eroded ridge <strong>of</strong> fine clay partially reforested with<br />

eucalyptus trees. At the northwest end <strong>of</strong> the ridge,<br />

a small gravely patch that looked like a possible<br />

relict deflation surface had a concentration <strong>of</strong> ten<br />

sherds <strong>of</strong> late Roman or early Medieval<br />

commonware and one fragment <strong>of</strong> Roman tile. The<br />

sherds all appear to be <strong>of</strong> the same vessel, a jar <strong>of</strong> a<br />

form known from other late Roman and early<br />

Medieval sites. This unusual concentration in such<br />

a small area, and especially the combination <strong>of</strong> tile<br />

with the remains <strong>of</strong> what may have been a whole<br />

vessel, suggests that this may have been a late<br />

Roman or early Medieval burial. If so, however, it<br />

has entirely eroded away, because the artifacts were<br />

all found resting on the apparent deflation surface.<br />

Site 28 (Papagallo, Area 101). Site, < 1<br />

ha? Prehistoric, Greek/Roman. Area 101 is located<br />

in an eroding badlands formation on the east slope<br />

<strong>of</strong> a clay ridge near the border between <strong>Bova</strong><br />

<strong>Marina</strong> and Cond<strong>of</strong>uri <strong>Marina</strong>. From the relatively<br />

small area <strong>of</strong> this slope that we investigated, we<br />

collected several fragments <strong>of</strong> prehistoric impasto<br />

and Greek or Roman commonware. The prehistoric<br />

pottery appeared to be later than the Neolithic<br />

assemblage at Umbro, but did not include any clear<br />

chronological diagnostics. The Greek or Roman<br />

pottery needs to be reexamined to determine its<br />

date. This area should be revisited to determine the<br />

extent <strong>of</strong> the site and to see whether contexts can be<br />

identified in the eroding banks from which the<br />

artifacts are eroding.<br />

Site 29 (Pisciotta F, Areas 8, 80, 81, 82,<br />

87). Scatter. Roman. The fields on the mildly<br />

sloping eastern side <strong>of</strong> the S. Pasquale valley, just<br />

to the east <strong>of</strong> the Torrente Turdari, contain a higher<br />

than usual frequency <strong>of</strong> Roman pottery. There does<br />

not appear to be a clear concentration anywhere<br />

16<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

within these areas, however, so it may all be a<br />

diffuse scatter associated with Site 2 or 13.<br />

Site 30 (Umbro D, Area 130). Site, ca. 0.6<br />

ha. Neolithic?, Bronze Age? This site, reported to<br />

us by S. Stranges, is located on the edge <strong>of</strong> an area<br />

at Umbro partially surveyed in 1998. It is a long,<br />

narrow scatter <strong>of</strong> artifacts along the southern edge<br />

<strong>of</strong> a plateau area at Umbro, where the slope begins<br />

to become broken and steeper. The finds are mostly<br />

nondiagnostic fragments <strong>of</strong> prehistoric pottery,<br />

along with one fragment <strong>of</strong> worked flint. The<br />

majority <strong>of</strong> the pottery resembles the Neolithic<br />

pottery from the excavations at Umbro in fabric,<br />

and one fragment appears to have impressed<br />

decoration, but some are somewhat finer and may<br />

be later. One <strong>of</strong> these may be a fragment <strong>of</strong> a hornshaped<br />

handle or some similar projecting element.<br />

Another rim, though <strong>of</strong> coarse fabric, appears to<br />

have the attachment for a strap handle projecting<br />

above the rim.<br />

Site 31 (Umbro E, Area 131). Site, ca.<br />

0.25 ha. Neolithic? This site is located on the<br />

central plateau at Umbro, in an area not surveyed<br />

previously. It is at the southern edge <strong>of</strong> the central<br />

plateau, at the top <strong>of</strong> some low cliffs overlooking<br />

the field containing Site 30. A scatter <strong>of</strong> prehistoric<br />

pottery (plus one piece <strong>of</strong> obsidian) was found<br />

along the top <strong>of</strong> the cliff. No useful diagnostic<br />

fragments were found, but the fabrics generally<br />

resemble the Neolithic pottery from the excavations<br />

at Umbro. That, together with the obsidian, makes a<br />

Neolithic date most likely for this site.<br />

Site 32 (Marcasita, Areas 136, 137). Site,<br />

ca. 0.8 ha. Neolithic?, Roman. This site is at the<br />

end <strong>of</strong> a low hill projecting into the eastern side <strong>of</strong><br />

the S. Pasquale valley, overlooking the floodplain.<br />

It is located directly across the river bottom from<br />

Site 14 (Panaghia). It comprises a fairly dense and<br />

well-defined scatter <strong>of</strong> brick, tile, and pottery in<br />

two clusters on the northeast and southwest sides <strong>of</strong><br />

the ridge. Some <strong>of</strong> the brick and tile are modern,<br />

related to an abandoned farmhouse just to the west,<br />

but some are clearly Roman. Roman imports<br />

include one fragment <strong>of</strong> Campanian (black sand)<br />

amphora and several nondiagnostic fragments <strong>of</strong><br />

African Red Slip ware, African plainware, and<br />

African amphora. The Roman assemblage does not<br />

contain anything that can be dated very closely but<br />

would support a date range from the 1st century AD<br />

(possibly somewhat earlier) to the 4th century AD<br />

or later. The size and density <strong>of</strong> the Roman<br />

component conform to what would be expected <strong>of</strong> a<br />

single farm site. The site also yielded a fair amount<br />

<strong>of</strong> prehistoric pottery, in fabrics resembling the<br />

Neolithic impasto and figulina from the<br />

excavations at Umbro. One fragment may have<br />

traces <strong>of</strong> impressed decoration. Thus, the


prehistoric assemblage may be <strong>of</strong> Neolithic date.<br />

The prehistoric pottery is less eroded than usual for<br />

surface assemblages in <strong>Bova</strong> <strong>Marina</strong>, which may<br />

mean that plowing has recently cut into or is<br />

currently cutting into the archaeological levels.<br />

Site 33 (Cromidi, Area 117). Site, < 1 ha?<br />

Greek, Roman? A small cluster <strong>of</strong> ancient pottery<br />

and tile, including one fragment <strong>of</strong> Greek blackgloss.<br />

It is located about 500 meters northeast <strong>of</strong><br />

the Neolithic site at Umbro, in the valley by the<br />

foot <strong>of</strong> the cliffs <strong>of</strong> M. Buccisa, near some springs.<br />

The surrounding parts <strong>of</strong> Area 117 also yielded two<br />

fragments <strong>of</strong> African Red Slip ware, as well as a<br />

fair amount <strong>of</strong> Greek and Roman tile. Thus, it is<br />

likely that there was a small occupation, possibly a<br />

farm or a seasonal fieldhouse, used in both Greek<br />

and Roman periods.<br />

Site 34 (Vunemo, Area 123). Site, < 0.1<br />

ha? Roman? A small, fairly level area <strong>of</strong> vines and<br />

fruit trees was the only part <strong>of</strong> Area 123 to yield<br />

artifacts in any quantity. A concentration <strong>of</strong><br />

artifacts in this one location included Greek or<br />

Roman coarse pottery and tile. No obvious<br />

chronological diagnostics were found, but the<br />

overall character <strong>of</strong> the assemblage may suggest an<br />

early Roman date.<br />

Site 35 (Cecilia, Area 104). Scatter.<br />

Roman. This consists <strong>of</strong> a small quantity <strong>of</strong> artifacts<br />

found during fieldwalking on the north face <strong>of</strong> a hill<br />

along the border between <strong>Bova</strong> <strong>Marina</strong> and Palizzi<br />

<strong>Marina</strong>. A large piece <strong>of</strong> Roman tegula was found<br />

on the slope below the crest, and a few scraps <strong>of</strong><br />

pottery, some <strong>of</strong> which may be Roman, were found<br />

near the top. A stone weight for an olive press,<br />

possibly Late Roman or later in date, was found<br />

built into the wall <strong>of</strong> a recent fieldhouse at the base<br />

<strong>of</strong> the hill to the north.<br />

Site 36 (Panaghia B, Area 143). Scatter.<br />

Roman. This area, on the terraces <strong>of</strong> the western<br />

side <strong>of</strong> the S. Pasquale valley, produced a diffuse<br />

scatter <strong>of</strong> Roman pottery and tile, none <strong>of</strong> it clearly<br />

diagnostic. There is no clear concentration here or<br />

in the slopes above this area, so although it could<br />

be a small, low-density site, it is at least as likely to<br />

be dense background scatter, possibly related to the<br />

Roman site at Panaghia (Site 14).<br />

Site 37 (Panaghia C, Area 145). Scatter.<br />

Greek, Roman? This area, on a slope overlooking<br />

the western side <strong>of</strong> the S. Pasquale valley, produced<br />

a diffuse scatter <strong>of</strong> pottery and tile, mostly modern<br />

(related to an abandoned farmhouse at the upper<br />

end <strong>of</strong> the area) but also including a few pieces <strong>of</strong><br />

Greek or Roman coarseware and tile. No<br />

concentration <strong>of</strong> ancient artifacts was observed, so<br />

this is likely to be background scatter related to a<br />

17<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

site located elsewhere, possibly the Roman site at<br />

Panaghia (Site 14).<br />

2.1.4. Discussion <strong>of</strong> survey results<br />

It may be useful at this point to give a<br />

brief review <strong>of</strong> what we have learned so far from<br />

our survey.<br />

We have not yet found any evidence <strong>of</strong><br />

Paleolithic activity. Given sporadic Paleolithic<br />

finds elsewhere in the area (at Torre Mozza and<br />

Gunì in Palizzi <strong>Marina</strong>, for instance; S. Stranges<br />

and L. Saccà, pers. comm.) it is unlikely that <strong>Bova</strong><br />

<strong>Marina</strong> was simply uninhabited before the<br />

Neolithic. Paleolithic settlements are likely to be<br />

archaeologically unobtrusive, and coastal sites are<br />

probably now under water due to rising sea levels at<br />

the beginning <strong>of</strong> the Holocene. However, the lack<br />

<strong>of</strong> Paleolithic sites may also be the result <strong>of</strong><br />

geomorphological processes having eliminated<br />

older ground surfaces in the study area. It is also<br />

possible that we have failed to recognize them<br />

adequately: our survey has recovered very few<br />

chipped stone artifacts so far, possibly due to our<br />

reliance on inexperienced crews. As the number <strong>of</strong><br />

later prehistoric sites grows, our understanding <strong>of</strong><br />

site location improves.<br />

Neolithic sites seem to occur in a wide<br />

variety <strong>of</strong> settings, including coastal valley terrace<br />

(Site 1), low hill overlooking a river valley (Site<br />

32), and inland plateau (Sites 9 and 31). Later<br />

prehistoric periods are more difficult to distinguish<br />

(see discussion <strong>of</strong> ceramic chronology above), but a<br />

similar diversity is likely.<br />

In the Classical periods, Greek settlement<br />

includes a large village (Site 12) as well as small<br />

farmsteads (Sites 18 and 33). As with the<br />

prehistoric sites, however, most are located on high<br />

ground, although some exceptions occur (Site 33).<br />

In the Roman period for the first time numerous<br />

sites are known from valley locations (Sites 2, 6,<br />

14, 22, and 26), suggesting a shift toward more<br />

intensive agricultural use <strong>of</strong> valley bottoms. The<br />

ceramic evidence for the Greek and Roman periods<br />

shows two peaks: one in late Archaic to Classical<br />

times (6th to 5th centuries BC) and one in late<br />

imperial times (3rd to 5th centuries AD). Late<br />

ancient settlement seems to continue in diminishing<br />

abundance but in the same locations as Roman<br />

settlement, to the 8th century or so (Site 22). After<br />

that, medieval settlement remains obscure, but by<br />

the 19th century abundant rural settlement again<br />

existed throughout <strong>Bova</strong> <strong>Marina</strong>, like in Greek and<br />

Roman times.<br />

In relation to the goals we had set for the<br />

1999 season, we made reasonable progress given


the constraints <strong>of</strong> such a short field season. The<br />

detailed evidence from Mazza is not only likely to<br />

be important for understanding Greek settlement<br />

patterns in the region, it also revealed significant<br />

prehistoric and Roman occupations, adding to our<br />

knowledge <strong>of</strong> those periods. The size <strong>of</strong> this site<br />

compared to others, however, and the anomalously<br />

high artifact density at the highest part <strong>of</strong> the hill,<br />

raises many questions about the internal structure <strong>of</strong><br />

the site. Geophysical prospection, for example by<br />

resistivity/conductivity or magnetometer survey,<br />

may help explain the differences in Greek and<br />

Roman settlement patterns, by allowing us to<br />

compare sites in terms <strong>of</strong> number and arrangement<br />

<strong>of</strong> recognizable structures.<br />

The recognition <strong>of</strong> a number <strong>of</strong> small,<br />

rural Greek sites in the study area, a considerable<br />

distance from any Greek town, is a valuable<br />

addition to the understanding <strong>of</strong> Greek colonization<br />

derived from the study <strong>of</strong> urban sites such as Locri<br />

and their immediate surroundings as at Metaponto.<br />

The early date <strong>of</strong> some <strong>of</strong> these sites is somewhat<br />

surprising; it indicates that Greek settlement spread<br />

very widely through the colonial areas within a few<br />

generations. What remains unclear is the function<br />

<strong>of</strong> these sites, and their relationship on the one hand<br />

to the towns and on the other to the native<br />

population. Excavation at a small, early Greek site<br />

such as Site 18 may provide evidence that can be<br />

compared to that from urban sites such as Locri in<br />

order to address such questions.<br />

The additional evidence from higher,<br />

interior locations (M. Rotonda, M. Vunemo) is a<br />

valuable complement to our evidence from Umbro.<br />

In particular, it shows two things. First, Umbro is<br />

unusual among the higher locations in <strong>Bova</strong><br />

<strong>Marina</strong>, in the amount <strong>of</strong> prehistoric settlement to<br />

be found there. Although both M. Rotonda and M.<br />

Vunemo produced some archaeological evidence,<br />

nothing like the cluster <strong>of</strong> prehistoric sites at<br />

Umbro was found. Second, even in highland areas<br />

that appear fairly barren at present, Greek and<br />

Roman settlement and activity were present to some<br />

degree. This suggests a problem with the present<br />

definition <strong>of</strong> our survey. In early modern times,<br />

<strong>Bova</strong> Superiore and <strong>Bova</strong> <strong>Marina</strong> formed a single<br />

unit, and it is likely that agricultural practices <strong>of</strong> the<br />

time made use <strong>of</strong> land at all elevations within this<br />

extended territory. There is no reason to suppose<br />

that the prehistoric, Greek, and Roman periods<br />

were different; interior areas in <strong>Bova</strong> Superiore<br />

were probably used by the same people using parts<br />

<strong>of</strong> <strong>Bova</strong> <strong>Marina</strong>, for cultivation <strong>of</strong> grain or other<br />

crops in the lower parts, for haying or pasture at<br />

higher altitudes, and even higher places could have<br />

been used for other purposes such as hunting,<br />

wood-cutting, charcoal-burning, or mineral<br />

extraction. Moreover, since we know that the<br />

18<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

principal focus <strong>of</strong> medieval settlement in the region<br />

was the interior town <strong>of</strong> <strong>Bova</strong>, the apparent absence<br />

<strong>of</strong> medieval settlement from our survey may be a<br />

result <strong>of</strong> working too near the coast. For these<br />

reasons, a survey concentrated entirely on the lower<br />

areas within the borders <strong>of</strong> modern <strong>Bova</strong> <strong>Marina</strong> is<br />

incomplete, and runs the danger <strong>of</strong> missing<br />

important parts <strong>of</strong> past settlement systems. This<br />

danger is greatest for the warmer periods <strong>of</strong> the<br />

past, when arable cultivation could have extended<br />

higher than in recent times; such periods may<br />

include important parts <strong>of</strong> the Bronze Age and the<br />

Middle Ages.<br />

The problems with the chronological<br />

sequence remain worrisome. It appears likely that a<br />

combination <strong>of</strong> continued survey and statistical<br />

seriation <strong>of</strong> our assemblages will improve the<br />

resolution for the Greek and Roman periods,<br />

already the best-known part <strong>of</strong> the sequence.<br />

Expanding the survey to higher areas near <strong>Bova</strong><br />

Superiore or Amendolea Vecchia may improve our<br />

knowledge <strong>of</strong> the medieval part <strong>of</strong> the sequence.<br />

The problems with the prehistoric sequence will<br />

require a different approach, however. Apart from<br />

chance finds <strong>of</strong> freshly disturbed prehistoric sites,<br />

an unlikely occurrence, survey is not likely to<br />

produce assemblages in good enough condition to<br />

allow systematic comparison <strong>of</strong> the attributes used<br />

in traditional classifications with those more readily<br />

visible in our surface collections, such as paste,<br />

surface treatment, and firing. Surface collections<br />

also lack secure archaeological contexts, so one<br />

does not know whether they represent an umixed<br />

assemblage and one cannot associate them with<br />

absolute dates. What may be required, instead, is a<br />

program <strong>of</strong> small test excavations at selected<br />

prehistoric sites, to obtain samples <strong>of</strong> less eroded<br />

artifacts from better-defined contexts, possibly in<br />

association with materials suitable for radiocarbon<br />

dating. Such excavations, carefully targeted at sites<br />

likely to have intact deposits <strong>of</strong> enigmatic date,<br />

would not only enable the excavated sites to be<br />

dated, but would improve the precision <strong>of</strong> the<br />

dating <strong>of</strong> our survey in general.<br />

Finally, we were asked by the<br />

Soprintendenza Archeologica della Calabria to<br />

watch for archaeological sites at imminent risk <strong>of</strong><br />

destruction due to human activity. While numerous<br />

sites are subject to destructive geomorphological<br />

processes such as erosion and human practices such<br />

as plowing, we observed only two sites actually<br />

threatened by active construction. One <strong>of</strong> these is<br />

Panaghia, an inland Roman villa or farmstead site<br />

where digging foundations for new houses had<br />

destroyed part <strong>of</strong> the site already. Construction has<br />

now been stopped at Panaghia following legal<br />

action by the Soprintendenza. The other site is<br />

Siderone/ Amigdala, located beneath the SS106


overpass on the eastern side <strong>of</strong> the modern town <strong>of</strong><br />

<strong>Bova</strong> <strong>Marina</strong>. This Roman villa or village site has<br />

been largely covered by modern building, and parts<br />

still preserved in gardens are currently being built<br />

upon. Again, the Soprintendenza has taken legal<br />

action to halt building and has also begun rescue<br />

excavations.<br />

2.2. Intensive Collections at Mazza (Lin<br />

Foxhall)<br />

Controlled surface collections were<br />

carried out over four days on site 12, Mazza, to<br />

determine the geographical limits, the<br />

chronological range and patterning, and the internal<br />

spatial organisation <strong>of</strong> this large and complex<br />

historic-period site.<br />

A baseline was established along the fence<br />

near the summit <strong>of</strong> the site. Using a compass, tapes<br />

and ranging rods, a grid oriented along cardinal<br />

points was plotted over the accessible areas <strong>of</strong> the<br />

site with stations marked with chaining pins every<br />

30 m. Each station was assigned a number and a<br />

letter designating its position within the grid, and<br />

each was then used as the centre point for a circle<br />

10m 2 in area. Circles were demarcated using two<br />

attached to a piece <strong>of</strong> string 1.78 m in length. One<br />

nail was held at the station point while the other<br />

was used to inscribe the circumference <strong>of</strong> a circle<br />

with a 1.78 m radius. All artefacts were collected<br />

from within each <strong>of</strong> these areas. Part <strong>of</strong> the site,<br />

near the (probably) early modern-modern structures<br />

previously identified as a ‘temple’, is fenced <strong>of</strong>f so<br />

it was not possible to include this area in our<br />

intensive systematic survey. However, the intensive<br />

survey area did cover the section <strong>of</strong> the site with<br />

modern (19th century?) farmhouse buildings, just<br />

south <strong>of</strong> the summit. A total <strong>of</strong> 105 small<br />

collections were made (Figure 3, Figure 4).<br />

Spatial data is summarized in a set <strong>of</strong><br />

distribution maps <strong>of</strong> selected artifact categories.<br />

Each map depicts the number <strong>of</strong> fragments <strong>of</strong> that<br />

category collected at each location; collection units<br />

with no items <strong>of</strong> that category are left blank. These<br />

numbers are based on preliminary sorting in 1999<br />

and may be subject to revision. A few problems<br />

with the data which are relevant to reading these<br />

maps should be noted. First, one bag, from<br />

collection H8, became separated from its contents<br />

while the finds were being washed. These artifacts<br />

presumably became mixed with those <strong>of</strong> one or<br />

more other bags. Second, an assortment <strong>of</strong> pottery<br />

and tile became mixed with the contents <strong>of</strong> a<br />

different bag, which originally contained samples<br />

<strong>of</strong> metal-smelting debris. This group <strong>of</strong> pottery and<br />

tile fragments is shown in the circle in the lower<br />

part <strong>of</strong> each map. The quantity and types resemble<br />

19<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

those <strong>of</strong> the units around H8, but it is not known<br />

whether these are in fact some or all <strong>of</strong> the missing<br />

finds. Finally, several units (L6 to L12) were<br />

incorrectly placed too far to the east. These were<br />

redone once the error was noted, but the initial<br />

mistaken collections are included on the maps<br />

between K and L.<br />

Preliminary analysis <strong>of</strong> the material<br />

suggests that temporal and spatial patterns are<br />

discernable in the controlled surface collections.<br />

Very large amounts <strong>of</strong> ancient ro<strong>of</strong> tile were found<br />

at the highest part <strong>of</strong> the site, smaller quantities on<br />

the rest <strong>of</strong> the upper areas and the SE slopes, and<br />

much less around the edges. Natural rock outcrops<br />

in the central portion <strong>of</strong> the site S and SW <strong>of</strong> the<br />

summit (Area 151), look similar to those used for<br />

house foundations at Locri Epizephyrii, and might<br />

perhaps have served the same purpose. The<br />

northern slopes <strong>of</strong> the site showed low level artefact<br />

scatter almost to the gully. On the southern and<br />

southeastern slopes artifact scatter continued as far<br />

as we could survey to the edges <strong>of</strong> the steep slopes,<br />

though density levels decreased dramatically<br />

toward the edges. The southern end <strong>of</strong> the site<br />

appears to have a concentration <strong>of</strong> smelting/metal<br />

working remains. Burnt daub and furnace slag were<br />

collected and a large piece <strong>of</strong> probably furnace slag<br />

too large to collect was photographed. Slag samples<br />

were taken to Leicester University for analysis.<br />

Ceramic data has not yet been fully<br />

analysed, but preliminary results suggest a<br />

preponderance among the Greek fine-wares <strong>of</strong> cups<br />

and krater-like shapes. Pottery is most abundant on<br />

the eastern, southern, and southwestern parts <strong>of</strong> the<br />

plateau, as well as on the southeastern terrace<br />

below; an area <strong>of</strong> at least 8 hectares in all. The<br />

artifact scatter is variable in density, but the only<br />

major anomaly is the dense concentration <strong>of</strong> tile at<br />

the summit. Greek pottery, as represented by blackgloss<br />

wares, occurs throughout this area. Roman<br />

pottery, as represented by African imports, occurs<br />

mostly in the eastern and southeastern parts <strong>of</strong> the<br />

site. Prehistoric pottery is most abundant in the<br />

southern and southwestern parts <strong>of</strong> the site, which<br />

is also the area where the metal-smelting evidence<br />

occurred.<br />

To conclude, intensive collection at Mazza<br />

has showed that there is both chronological<br />

patterning within the site, with Greek remains<br />

everywhere and an unsuspected Roman component<br />

concentrated near the highest part <strong>of</strong> the site. It has<br />

also shown that there is functional differentiation<br />

within the site, with possible residential and<br />

industrial areas identifiable. Further analysis <strong>of</strong> this<br />

data should clarify these patterns.


2.3. Geological Reconnaissance<br />

Prehistoric people made use <strong>of</strong> a range <strong>of</strong><br />

geological raw materials; archaeologically<br />

demonstrable examples include clay, flint, obsidian,<br />

fossil shark teeth, hard stones for polished axes, and<br />

stone for querns, mills and other ground stone tools.<br />

The sources <strong>of</strong> some <strong>of</strong> these are known. Obsidian<br />

used at Umbro generally came from Lipari (R.<br />

Tykot, pers. comm.) and some grey flint came from<br />

outcrops in the Saracena area <strong>of</strong> Cond<strong>of</strong>uri <strong>Marina</strong>.<br />

The polished stone axe found at Umbro in 1999<br />

(see below) was made from amphibolite from the<br />

Sila (P. Barrier, pers. comm.), and the miniature<br />

axe replica was made from a s<strong>of</strong>ter phyllite schist<br />

available locally. The source <strong>of</strong> other materials<br />

remains unknown. Petrographic analysis has shown<br />

that pots probably made at Umbro were not made<br />

from clay from the two sources nearest to the site<br />

(see below), and we have little idea where the red,<br />

brown, and yellow flint used at Umbro came from.<br />

Hard stone for making axes has been the<br />

object <strong>of</strong> a number <strong>of</strong> sourcing studies. Throughout<br />

Southern Italy and Sicily, prehistoric people used<br />

granite, diorites, serpentine and other greenstones,<br />

amphibolite, and other stones for axes, adzes, and<br />

similar tools. Often greenstones transported from<br />

long distances were used for smaller or finer tools,<br />

particularly amulets, while amphibolite and diorite<br />

were used for less fine axes and more resistant and<br />

commoner stones such as granite were used large<br />

pebble tools (Leighton 1999; O’Hare 1990). These<br />

studies make clear both that hard stones were<br />

transported over considerable distances, with some<br />

coming ultimately from the Alps, and that materials<br />

were <strong>of</strong>ten carefully matched with a tool’s intended<br />

form and use. It is also probable that the trade in<br />

stone and axes not only carried on to supply useful<br />

tools but was socially and symbolically important<br />

as well.<br />

With these considerations in mind, we<br />

decided to investigate the availability <strong>of</strong> hard stones<br />

in the <strong>Bova</strong> <strong>Marina</strong> region. If appropriate raw<br />

materials were available, could they have been an<br />

economic resource for local axe production or for<br />

trade? If raw materials were not available locally,<br />

were they available inland in the highlands <strong>of</strong><br />

Aspromonte? If so, what were the implications for<br />

trade? If raw materials were available and were not<br />

used, why would this have been so?<br />

The geology <strong>of</strong> Aspromonte is highly<br />

varied, but the potential sources <strong>of</strong> polishable hard<br />

stone seem limited. The territory <strong>of</strong> <strong>Bova</strong> <strong>Marina</strong><br />

itself, like much <strong>of</strong> the coastal strip, consists <strong>of</strong><br />

harder and s<strong>of</strong>ter sedimentary stones (limestones,<br />

sandstones, clays, shales) and <strong>of</strong> metamorphic<br />

20<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

rocks including predominantly phyllites, schists and<br />

gneisses. The latter are <strong>of</strong>ten too s<strong>of</strong>t or too riven<br />

by cleavage planes to be used for tools. Harder and<br />

finer grained metamorphic and igneous stones<br />

occur at higher altitudes 10-20 km inland; these<br />

include granites, diorites, and possibly other stones<br />

(Servizio Geologico dell’Italia 1970; 1976).<br />

Given this, hard stones could have been<br />

available locally in two ways. First, fragments<br />

could have washed down from the highlands in<br />

river valleys. Secondly, a band <strong>of</strong> relatively s<strong>of</strong>t<br />

sandstone conglomerate stretches across the<br />

territory at about 3-6 km inland. As road cuts in the<br />

Fiumara di Amendolea valley, Monte Bucissa, and<br />

alnog the <strong>Bova</strong> Superiore road show, this<br />

conglomerate contains strata dense with cobbles <strong>of</strong><br />

varied stones which must have originally come<br />

from older strata higher inland. A cursory<br />

inspection on Monte Bucissa showed that most <strong>of</strong><br />

these were granites and gneisses, but they also<br />

contain some unidentified close-grained stones<br />

which may have been suitable for tool-making. The<br />

only crystalline material was a coarse-grained and<br />

fault-ridden quartzite. These conglomerate strata<br />

would have provided another potential source <strong>of</strong><br />

stone for tools.<br />

The logical strategy for assessing raw<br />

material availability, therefore, was through<br />

surveying river valleys where both recently<br />

transported stone and cobbles eroding from<br />

conglomerates would be available. In addition to<br />

recording raw materials present, we hoped to<br />

collect data on how much stone <strong>of</strong> each type was<br />

available, to assess the “richness” <strong>of</strong> potential<br />

source zones quantitatively. As a pilot experiment,<br />

we recorded stone resources at tweve locations<br />

along the Fiumara di Amendolea. These were<br />

spaced approximately a kilometer apart, extending<br />

inland from the coast to above the ancient castle <strong>of</strong><br />

Amendolea. At each point, the crew laid out a<br />

circle enclosing ten square meters, and collected all<br />

cobbles larger than 15 cm long by 10 cm high by 5<br />

cm wide. They then matched these to a type<br />

collection they carried with them and tallied how<br />

many cobbles <strong>of</strong> each type <strong>of</strong> stone were present.<br />

When a new type <strong>of</strong> stone was found, a fragment<br />

was added to the collection. Although cumbersome,<br />

this procedure allowed survey to be carried out<br />

without a trained geologist present on the crew and<br />

without collecting and transporting hundreds <strong>of</strong><br />

kilos <strong>of</strong> cobbles. The type collection was retained<br />

for later identification <strong>of</strong> samples by a trained<br />

petrologist.<br />

Results are pending at this time. However,<br />

it is clear from this exercise that plentiful cobbles<br />

<strong>of</strong> an appropriate size and shape are available in the<br />

major river valleys, and at least some <strong>of</strong> them are


hard, potentially usable stones. It is also clear that<br />

different points in river valleys varied in their raw<br />

material “richness;” this may have had implications<br />

for landscape use and social rights <strong>of</strong> access.<br />

2.4. GIS Analysis <strong>of</strong> Prehistoric Land Use<br />

(Doortje Van Hove)<br />

Much archaeological research focuses<br />

upon human settlements, ceremonial sites, and the<br />

material that is found within them rather than on the<br />

spaces in between. This is because ‘<strong>of</strong>f-site<br />

archaeology’ (Foley 1981) provides poorly studied<br />

material to work with and immense methodological<br />

difficulties. However, humans live in a world<br />

composed <strong>of</strong> many different kinds <strong>of</strong> places, so<br />

restricting our view <strong>of</strong> prehistoric people to discrete<br />

sites will inevitably and unnecessarily narrow our<br />

understanding <strong>of</strong> how they lived. In emphasising<br />

the fuller landscape, this research highlights human<br />

usage <strong>of</strong> land, one <strong>of</strong> the key means by which<br />

humans categorise and lay claim to their<br />

surroundings.<br />

This research is aimed at understanding<br />

how people used the landscape in prehistoric<br />

Southern Italy through the use <strong>of</strong> computer GIS<br />

(Geographical Information Systems). The overall<br />

goal is to create a simulated environment<br />

representing one possible situation in which<br />

Southern Italians could have lived throughout the<br />

different periods <strong>of</strong> human occupation <strong>of</strong> the area.<br />

Particular attention is paid to land use and resource<br />

exploitation within the possible economies <strong>of</strong> each<br />

period, and the possible cultural categorisation and<br />

social use <strong>of</strong> <strong>of</strong>f-site areas.<br />

To investigate the scope <strong>of</strong> the research, a<br />

pilot study was done which focussed on a very<br />

limited area during the Neolithic period. This<br />

preliminary research has drawn together elements<br />

<strong>of</strong> archaeology, human psychology, geography and<br />

simulation modelling to provide novel insights into<br />

human-landscape relations. The research simulated<br />

environments and human land uses around three<br />

Neolithic sites: Castello <strong>Bova</strong> Superiore, Umbro<br />

and San Pasquale (see Figure 5). The quantities <strong>of</strong><br />

archaeological and environmental data and the<br />

spatial dimensions <strong>of</strong> the area examined were very<br />

limited. This prototype allowed developments to<br />

be made in the techniques and interpretative<br />

frameworks, before applying the methodology on a<br />

larger scale.<br />

GIS modelling, based upon relevant data<br />

from IGM (Istituto Geografico Militare)<br />

topographic and geological maps, and limited field<br />

survey site distributions from the <strong>Bova</strong> <strong>Marina</strong><br />

<strong>Archaeological</strong> <strong>Project</strong>, was used. Environmental<br />

relations were modelled using current and historical<br />

21<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

Mediterranean geographical and ecological<br />

information, and comparative studies from<br />

Mediterranean prehistory. In the broader area<br />

around Umbro, incorporation <strong>of</strong> different kinds <strong>of</strong><br />

environments allowed us to model the travel<br />

distances and sizes <strong>of</strong> specific activity-zones <strong>of</strong> the<br />

Neolithic cultures in the region.<br />

As basis for the analysis, topographic and<br />

geological maps were digitised. The base map used<br />

was the IGM 1:25000 map quadrant for <strong>Bova</strong><br />

<strong>Marina</strong>. Eight categories <strong>of</strong> environment were<br />

defined on the basis <strong>of</strong> the elevation, slope, river<br />

and geology maps. For this purpose, parameters<br />

were established on the basis <strong>of</strong> literature,<br />

assumptions and observations <strong>of</strong> the terrain in order<br />

to represent and evaluate the present and past<br />

landscape. General map algebra was used, which<br />

involves calculations with different categories <strong>of</strong><br />

various maps to generate a new map, which can<br />

give novel insights into a newly generated<br />

environment. The eight different environments are<br />

coastal plain, river channel, river valley, clay<br />

terrace, sedimentary plateau, sedimentary hill,<br />

metamorphic plateau and metamorphic hill.<br />

To evaluate the potential usefulness <strong>of</strong><br />

each zone on the map and its possible utility for<br />

humans, a combination <strong>of</strong> yield values and<br />

accessibility had to be attributed to each<br />

environmental niche. Thus, the environment<br />

reconstruction map was reclassified to hold<br />

estimated resource yield values. These valued were<br />

estimated based on the usefulness <strong>of</strong> each<br />

environmental zone to a specific kind <strong>of</strong> Neolithic<br />

economy. Four kinds <strong>of</strong> possible economy in the<br />

Neolithic have been selected for the analysis:<br />

foraging, mixed foraging-farming, subsistence<br />

horticulture and intensive pastoralism.<br />

In assessing how people will use their<br />

environment, not only the qualitative yield but also<br />

the accessibility <strong>of</strong> the area plays a significant role.<br />

For each site, a cost-surface model was built to<br />

assess the possible access to areas around the<br />

different sites. Therefore, a friction surface was<br />

created based upon a valid and plausible<br />

assumption <strong>of</strong> the cost traversing a sloping<br />

landscape. The area is quite hilly, and therefore it<br />

was considered that a given distance will be<br />

traversed with different time and effort costs on<br />

level or on sloping ground. In addition, walking<br />

uphill or downhill will affect time-intervals as well.<br />

It becomes increasingly more difficult and time<br />

consuming as the slope increases. To understand<br />

the values <strong>of</strong> the cost-surface model, a calibration<br />

was made, based on a plausible assumption about<br />

the landscape. Five classes <strong>of</strong> ‘distance’ or ‘cost’<br />

were defined to calculate what it takes to get from


each site to the different environmental zones<br />

defined earlier.<br />

Combining yield and accessibility values,<br />

three kinds <strong>of</strong> dynamic simulations were run on a<br />

fixed population-size (50 people) for purposes <strong>of</strong><br />

simplicity. Each simulation can be considered as a<br />

sequence <strong>of</strong> events. The first simulation shows how<br />

far people need to travel before reaching enough<br />

land for their particular economies. The assessment<br />

<strong>of</strong> the size <strong>of</strong> land needed per population and<br />

economy was based on literature, especially the<br />

work <strong>of</strong> Susan Gregg (Gregg 1988). The value <strong>of</strong><br />

specific land use and its land size desired will<br />

depend on the economy used. Farmland will be<br />

more important in the horticulturalist and mixed<br />

farming-foraging economies than in the pastoralist<br />

and forager ones. For foraging a larger territory is<br />

required, although it does not have the same impact<br />

on the land. For herding, a slightly larger terrain<br />

will also be necessary as animal husbandry is more<br />

mobile than farming. This will have an effect on<br />

the time and distance needed to cover to achieve<br />

the suitable land size for the amount <strong>of</strong> people<br />

concerned.<br />

A second kind <strong>of</strong> simulation is<br />

vegetational growth. One starts from a set <strong>of</strong> empty<br />

maps, which all grow a particular kind <strong>of</strong><br />

vegetation during a fixed time period. After the<br />

vegetation has iterated for a particular amount <strong>of</strong><br />

time, the simulation stops and assumptions can be<br />

made about when a vegetation type reaches its<br />

climax state as each type grows at its own rate. It<br />

was assumed that wet plants will grow faster than<br />

plants on heavy soil, depending on the soil texture,<br />

temperature and precipitation. Beach vegetation is<br />

conditioned to withstand a lot <strong>of</strong> extreme climatic<br />

and edaphic conditions and will grow at an average<br />

rate. Varied plants were given a rate <strong>of</strong> growth in<br />

between the other types <strong>of</strong> vegetation.<br />

For the purpose <strong>of</strong> simplicity, both<br />

previous simulations were combined. First, the<br />

natural vegetation grows for a fixed amount <strong>of</strong><br />

years until it is mature enough to let people use it<br />

for their specific purposes. Humans then use the<br />

land for farming, herding or foraging for a certain<br />

period, after which these are abandoned and the<br />

natural vegetation can grow back slowly. Because<br />

<strong>of</strong> the different impacts that foraging, farming and<br />

herding can have on the environment, different<br />

parameters were used in the regeneration <strong>of</strong> the<br />

vegetation. This is because land that has been used<br />

for herding and agriculture is a lot more degraded<br />

than when it was used for foraging and it will take<br />

longer for natural vegetation to grow back. Huntergatherers<br />

tend not to exhaust their environments in<br />

order to reserve the possibility <strong>of</strong> returning to them<br />

after several seasons. The result is a total <strong>of</strong> 36<br />

22<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

maps, defining land use per site, economy and<br />

specific type <strong>of</strong> impact on the environment. Some<br />

examples are displayed as Figure 5.<br />

When simulating these land use maps,<br />

some basic assumptions were used, which applied<br />

to the yield and accessibility. For people to use a<br />

particular landscape zone, the yield has to be<br />

sufficiently high in order to be selected. In<br />

addition, it is assumed that foraging people will<br />

travel further than those herding and farming.<br />

Therefore the cost assumptions are different per<br />

specific land use. This will influence the land use<br />

maps and is interesting in comparative studies. It<br />

was also moderately based upon literature about<br />

site catchment analysis (Higgs and Vita-Finzi 1972:<br />

27-36; Jarman, Bailey and Jarman 1982: 26-46).<br />

The analysis thus resulted in maps<br />

illustrating anthropogenic environmental impact<br />

and potential spatial dimensions <strong>of</strong> the cultural<br />

world that humans lived in. These results have two<br />

basic implications. One is economic and ecological:<br />

human landscape use and environmental impact can<br />

be examined in a systematic way, in contrast to<br />

isolated single site-based studies. The second is<br />

cultural: humans use the landscape, including ‘<strong>of</strong>fsite’<br />

areas for many other non-economic uses.<br />

Many <strong>of</strong> these uses are conditioned by the existent<br />

human use <strong>of</strong> those areas.<br />

In this regard, it is interesting to observe<br />

that sites are preferably ecotonally located,<br />

enabling the exploitation <strong>of</strong> two or even three<br />

resource zones: cultivation on a plateau, some<br />

pasture for grazing and a possibility for hunting and<br />

gathering in wooded alluvial margins.<br />

The primary result <strong>of</strong> the GIS study to date<br />

is to underline the importance <strong>of</strong> non-farming land<br />

uses for site distribution. Instead <strong>of</strong> the commonly<br />

held view that competition for farmland was <strong>of</strong><br />

greatest importance in farming societies, the<br />

simulation showed that even in these, human<br />

interaction was probably structured by non-farming<br />

uses such as the needs <strong>of</strong> foraging and pasture. The<br />

sizes <strong>of</strong> the different specific land uses such as for<br />

foraging, mixed foraging-farming, farming and<br />

herding show a different picture. In the region<br />

under consideration, farmland is stretched out in<br />

small patches around the sites, while foraging and<br />

herding zones are continuous and large. Neolithic<br />

groups were probably more mobile, using bigger<br />

territories and moving frequently within them, than<br />

is usually assumed (e.g. (Higgs and Vita-Finzi<br />

1972). This means that the distribution <strong>of</strong> sites,<br />

even in mountainous areas like Calabria, should<br />

not be limited to available farmland. Observing the<br />

contacts between the areas <strong>of</strong> human impact<br />

associated with different sites enables inference <strong>of</strong><br />

possible elements <strong>of</strong> South Italian social and


cultural behaviour. These might include exchange<br />

and competition.<br />

GIS study <strong>of</strong> prehistoric land use in<br />

Calabria will be continued as part <strong>of</strong> an ongoing<br />

PhD thesis. Future developments include<br />

comparisons with actual land use at each site, as<br />

indicated by archaeological data, which could<br />

further support studies <strong>of</strong> site potential and the<br />

development <strong>of</strong> an environmental and economical<br />

site-type model for Southern Italy. Other future<br />

possible adaptations <strong>of</strong> the analysis and approach<br />

will seek to highlight a more pluralistic view on<br />

Neolithic land use. These would implement a<br />

higher level <strong>of</strong> detail and environmental data such<br />

as water resources, soil types and economic<br />

resource needs (biomass, nutritional components <strong>of</strong><br />

animals and plants). An expanded site database<br />

would also be introduced. In addition, different<br />

parameters considering human resource needs can<br />

be implemented and compared to see whether they<br />

make a significant difference in land use around<br />

each <strong>of</strong> the analysed sites. Finally, ongoing research<br />

will address current problems representing the<br />

temporal dimension in modelling. In part, it will<br />

examine population shifts across the landscape in<br />

response to resource exhaustion and the differences<br />

<strong>of</strong> human land use during the whole period <strong>of</strong><br />

occupation <strong>of</strong> this area.<br />

23<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


3.1. Introduction: previous work, goals and<br />

methods for this season<br />

Umbro is one <strong>of</strong> several names for a<br />

calcareous sandstone and limestone plateau about a<br />

kilometer in diameter located along the unpaved<br />

<strong>Bova</strong> <strong>Marina</strong> – <strong>Bova</strong> Superiore road just south <strong>of</strong><br />

the border between the two comunes (Figure 7).<br />

The plateau consists <strong>of</strong> sloping fields between 360<br />

and 400 meters above sea level, surrounded on all<br />

sides by sporadic, interrupted cliffs 10-20 meters<br />

high. On some sides, the cliffs were probably<br />

penetrated by shallow caves in the past; in other<br />

zones they afford vertical sheltering walls. To the<br />

east and south underlying, impermeable clay beds<br />

are exposed, and there may have been springs at the<br />

foot <strong>of</strong> the cliffs here in the past.<br />

The prehistoric site at Umbro is located<br />

along the east margin <strong>of</strong> the plateau (Figure 6,<br />

Figure 7). It has been the subject <strong>of</strong> about a decade<br />

<strong>of</strong> investigation before the work described here (see<br />

Robb 1997; 1998; Stranges 1992; Stranges and<br />

Saccà 1994; Tinè 1992). Stranges, Saccà and coworkers<br />

have made periodic informal surface<br />

collections since their original discovery <strong>of</strong> the<br />

Neolithic site in the early 1990s. Tinè conducted a<br />

walkover <strong>of</strong> the site in 1992, and dug a shallow,<br />

.5m square test pit, unfortunately in a sterile area <strong>of</strong><br />

the site. The site was re-surveyed by the <strong>Bova</strong><br />

<strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> in 1997 and 1998.<br />

Finally, the only excavations at the site were<br />

conducted in 1998 by the <strong>Bova</strong> <strong>Marina</strong><br />

<strong>Archaeological</strong> <strong>Project</strong>. These excavations put a<br />

one by two meter sondage in the main Neolithic<br />

area (Trench 1), excavated a one by one meter<br />

sondage along the south slopes <strong>of</strong> the cliffs (Trench<br />

3) and on the valley floor just below Trench 1<br />

(Trench 5), and an area excavation atop the cliffs<br />

which revealed post-Neolithic, probably Bronze<br />

Age deposits.<br />

These researches established several basic<br />

facts about the prehistoric use <strong>of</strong> the site:<br />

• The general area <strong>of</strong> Umbro was occupied in the<br />

Neolithic and in at least one phase <strong>of</strong> the<br />

Bronze Age, possibly the Middle Bronze Age<br />

(Tinè 1992).<br />

• Occupation was patchy; for instance, the<br />

Neolithic area, though rich in archaeological<br />

materials, was restricted to a small area below<br />

the main cliffs at the site.<br />

• Human bone at the Neolithic site was actually<br />

<strong>of</strong> Late Roman date, though no other Roman<br />

3. THE EXCAVATIONS AT UMBRO<br />

24<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

remains to speak <strong>of</strong> were found here. The<br />

origins <strong>of</strong> this deposit remain a puzzle.<br />

• The main Neolithic area was dated<br />

radiometrically to the mid-6 th and mid-5 th<br />

millennia (both Stentinello periods). It was<br />

probably also occupied later in the Late<br />

Neolithic (Diana) and possibly in the Copper<br />

Age as well.<br />

• The depositional sequence at the main<br />

Neolithic area is complex and difficult to<br />

interpret, particularly in the upper levels; in<br />

contrast, other areas such as the Bronze Age<br />

area in Trench 4 typically have shallow and<br />

simple stratigraphies.<br />

We approached the 1999 excavations with<br />

a number <strong>of</strong> research questions in mind:<br />

(1) one basic goal was to clarify and<br />

elaborate the dating <strong>of</strong> Neolithic sequence in<br />

Trench 1.<br />

(2) the Trench 1 Neolithic sondage was<br />

undertaken primarily to explore the cultural<br />

sequence and the potential for further investigation.<br />

In 1999 we hoped to excavate a much broader area<br />

to yield substantial data on Neolithic economy,<br />

settlement and culture.<br />

(3) investigations in other parts <strong>of</strong> the site<br />

were intended to clarify the presence or absence <strong>of</strong><br />

any further Neolithic deposits, for instance on the<br />

clifftops above Trench 1 and the fields atop the<br />

plateau. They were also intended to provide further<br />

dating material for the post-Neolithic but poorly<br />

dated area located in Trench 4.<br />

The basic methods remained unchanged<br />

from those used in 1998 (see Robb 1998 for<br />

description). To summarize, all excavations were<br />

located within an overall site grid whose datum was<br />

<strong>of</strong>f the edge <strong>of</strong> the site to the east. For each trench a<br />

local trench datum was established from which<br />

depths could be measured and converted to<br />

absolute depths. For Trenches 1 and 7, all soil was<br />

removed in 10 cm arbitrary levels; records were<br />

kept <strong>of</strong> how these related to the sometimes clear,<br />

sometimes fuzzy natural soil units. Trench 6 was<br />

excavated in arbitrary 10 cm levels except for the<br />

removal <strong>of</strong> large rock fall along the east side <strong>of</strong> the<br />

trench, where this was impractical. All deposits<br />

removed were sifted in 5 mm screens. Finds were<br />

bagged separately for each level and 1x1 meter<br />

square. When features such as pits or postholes<br />

were encountered, they were excavated and finds<br />

bagged separately as separate contexts. Carbon


samples were photographed in situ and removed<br />

with metal tools only to foil envelopes. Flotation<br />

samples were also removed from selected levels as<br />

detailed below. Following excavation, all trenches<br />

were mapped, pr<strong>of</strong>iled and photographed.<br />

3.2. Description <strong>of</strong> 1999 Trenches<br />

3.2.1. Trench 1 (continued excavation)<br />

Trench 1 was located just below the cliffs<br />

in the densest area <strong>of</strong> Neolithic artifacts (Figure<br />

10). It was dug into sediments more than 2 meters<br />

deep, which had accumulated on top <strong>of</strong> large<br />

boulders which had fallen from the rock face before<br />

the Neolithic period, creating a small talus cone<br />

about 3 meters high at the base <strong>of</strong> the cliffs.<br />

Excavations in Trench 5, at the base <strong>of</strong> these<br />

boulders, in 1998 had established that these<br />

boulders lie directly upon sterile clay and thus<br />

predate any human occupation <strong>of</strong> the site.<br />

In 1999 we opened a long narrow trench<br />

extending six meters north-south by one meter eastwest.<br />

After excavating the top 10-20 centimeters in<br />

four <strong>of</strong> these six squares, we changed strategy and<br />

excavated one <strong>of</strong> them (-8n/-36e) and the adjacent<br />

square to the west (-8n/-37e) as a deep sondage.<br />

Thus at the beginning <strong>of</strong> the 1999 season Trench 1<br />

essentially consisted <strong>of</strong> a one by two meter<br />

sondage. The goal <strong>of</strong> the 1999 season was to finish<br />

excavating all six meters <strong>of</strong> the original north-south<br />

trench down to bedrock or sterile soil (Figure 8).<br />

By the end <strong>of</strong> the season, this goal was<br />

largely accomplished through work with trowels,<br />

handpicks and pick. The only major difficulty was<br />

the removal <strong>of</strong> excavated soil, since there was little<br />

unexcavated area atop the talus for a backdirt heap,<br />

while shipping soil down to the base <strong>of</strong> the slope<br />

with buckets and ropes was very hard work and<br />

unsafe. In the end, we devised a system by which<br />

soil was poured down a plastic pipe about 20 cm in<br />

diameter and three meters long which dumped it<br />

directly in a screen hung from a tripod. Finds were<br />

then passed back up to the trench to be bagged<br />

inside small plastic containers. This system may<br />

have resulted in slight micro-damage in transit to a<br />

few delicate finds, but allowed excavation to<br />

continue rapidly without undue crew fatigue. The<br />

problem <strong>of</strong> ultimately carrying the soil back up the<br />

three meter drop to the trench to fill the trench back<br />

in remains unresolved.<br />

In the northernmost two squares <strong>of</strong> the<br />

trench (-6n/-36e and –7n/-36e), we stopped at the<br />

end <strong>of</strong> the season in sediments which, though still<br />

including sporadic Neolithic material, were<br />

increasingly sterile. In –8n/-36e, most <strong>of</strong> the<br />

sediments had been removed in the previous year’s<br />

25<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

work, but we excavated a further 10-20 cm. At the<br />

bottom, about half the square consisted <strong>of</strong> bedrock<br />

and open, air-filled fissures were gaping between<br />

buried boulders along the east side <strong>of</strong> the pit.<br />

Deposits in the next three squares to the south (-<br />

9n/-36e, -10n) were relatively shallower, as the<br />

underlying bedrock sloped upward. Excavation in<br />

these squares finished when bedrock was reached.<br />

We also excavated another 30-40 cm in the<br />

westward extension square (-8n/-37e) which<br />

reached a depth <strong>of</strong> about 280 cm below the trench<br />

datum.<br />

Unexcavated deposits remain at the base<br />

<strong>of</strong> the northern half <strong>of</strong> the square, at the bottom <strong>of</strong><br />

the westward extension, and below a large boulder<br />

just northeast <strong>of</strong> the trench. It is clear that the<br />

underlying boulders slope sharply downwards as<br />

they approached the cliff, and the ultimate depth <strong>of</strong><br />

deposits close to the cliff face remains unknown.<br />

The stratigraphy was described in terms <strong>of</strong><br />

five strata, as in the 1998 sondage (Figure 9; Table<br />

1). These were more or less found throughout the<br />

new, larger excavated area, with some ambiguity at<br />

times.<br />

Stratum I. This stratum consisted <strong>of</strong> a<br />

layer <strong>of</strong> topsoil 5-10 cm thick, with an irregular<br />

conformation and extensively disturbed by erosion,<br />

soil washing in, and vegetation. A loam containing<br />

sand, silt and clay, it was a light brownish color<br />

(Munsell color 10YR 4/2 dry). It contained cultural<br />

finds from all periods within the Neolithic.<br />

Stratum II. This stratum consisted <strong>of</strong> the<br />

same basic sediment as Stratum I, but was less<br />

disturbed with less vegetation. It was slightly<br />

sandier in places, probably due to the<br />

decomposition <strong>of</strong> sandstone slabs, and consisted <strong>of</strong><br />

a mixed sandy loam (Munsell color 10YR 5/3 dry).<br />

It contained some rock fall, particularly small rocks<br />

20-40 mm in diameter, and much penetration by<br />

small roots. Stratum II varied greatly in thickness,<br />

from about 20 cm up to about 60 cm. While its top<br />

surface was irregular, its bottom surface was far<br />

closer to horizontal. It contained cultural finds from<br />

all periods within the Neolithic as well as almost all<br />

the possible Copper Age sherds.<br />

Stratum III. This stratum consisted <strong>of</strong><br />

light, yellowish, gritty sediments similar to those in<br />

Stratum II (sandy loam, Munsell color 10YR 5/4<br />

dry), and in places the distinction between the two<br />

strata was not particularly clear. It contained more<br />

rock fall <strong>of</strong> all sizes. Rocks generally lay roughly<br />

horizontal, suggesting gradual rather than<br />

catastrophic accumulation. The differences between<br />

Statum II and Stratum III may be due in part to less<br />

root penetration in Stratum III, which might leave<br />

larger, undecomposed rocks while Stratum II was


slightly sandier. The north end <strong>of</strong> the trench<br />

contained considerable rock fall in this stratum with<br />

accompanying sandy lenses. Stratum III was a<br />

relatively horizontal level between 30 and 50 cm<br />

thick. It contained both Stentinello wares and<br />

abundant Diana wares (see below for further<br />

discussion).<br />

Stratum IV. This stratum was clearly<br />

distinguishable from Stratum III in section. It was a<br />

denser, browner, and more compact sandy clay with<br />

much less rock fall in it (Munsell color 10YR 4/4,<br />

dry). The rocks in this layer lay more or less<br />

horizontal. In the northern end <strong>of</strong> the trench,<br />

Stratum IV formed a homogeneous layer between<br />

30 and 40 cm thick. In the southern half <strong>of</strong> the<br />

trench, Stratum IV was less thick and overlay rising<br />

bedrock. It contained Stentinello pottery.<br />

Stratum V. This stratum made up the<br />

bottom 20-30 cm <strong>of</strong> excavated deposits in the<br />

northern end <strong>of</strong> the square. Although bedrock<br />

underlay this sediment directly in –8n/-36e, its<br />

depth at the northern end <strong>of</strong> the trench is unknown.<br />

It consisted <strong>of</strong> a dark brown (Munsell color 10YR<br />

4/4, dry) dense, compact, clayey sediment with<br />

little rock fall in it. It contained small flecks <strong>of</strong><br />

charcoal and disintegrated animal bone, resembling<br />

midden deposits. It contained Stentinello pottery.<br />

This definition <strong>of</strong> the stratigraphy is based<br />

upon the six-meter north-south strip <strong>of</strong> squares in –<br />

36e. In the westward extension square (-8n/-37e) all<br />

strata appear to slope downward sharply as one<br />

approaches the cliff face. This clearly has<br />

implications for both dating the site and<br />

interpreting the living area available at any one<br />

time. Clarifying the depth <strong>of</strong> the stratigraphy close<br />

to the cliff wall is a major goal <strong>of</strong> future excavation<br />

seasons.<br />

Nowhere in the excavated area did we<br />

encounter clear evidence <strong>of</strong> structures. Several<br />

faint, possible pits or post-holes all turned<br />

unconvincing upon full excavation, and may easily<br />

have been due to ancient disturbances by rodents or<br />

roots. Concentrations <strong>of</strong> small, dispersed, lightly<br />

fired fragments <strong>of</strong> daub were noted in certain<br />

levels, notably in Stratum III. Here the fragments<br />

consisted <strong>of</strong> flattish pieces ca. 5 cm thick, smoothed<br />

on both sides and with occasional impressions <strong>of</strong><br />

sticks up to 3 cm in diameter. These may be<br />

evidence <strong>of</strong> a house, wall, or shed; alternatively,<br />

they may be from a non-habitation structure such as<br />

a wall or daub-lined pit. When found, they were<br />

scattered in the soil at a consistent depth but<br />

without contextual position or orientation.<br />

Likewise, few features were found, and<br />

none which were unequivocally interpretible.<br />

Several large sand lenses were encountered. One, in<br />

26<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

the far southwestern corner <strong>of</strong> the pit, contained<br />

almost pure brownish sand ca. 25 cm deep in a<br />

neatly delimited area about 25 cm in area. The<br />

source <strong>of</strong> such sand is not clear as it does not occur<br />

on the site, nor in most geological contexts. Hence<br />

it may have been transported to the site for a<br />

specific purpose. In contrast, several sand lenses in<br />

–9n/-36e, -8n/-36e, and –6n/-36e were clearly<br />

derived from decomposing bedrock. They consisted<br />

<strong>of</strong> yellowish sand in broad, flat, generally<br />

horizontal lenses, diffusely bounded. The sand was<br />

very similar to that demonstrably freed from<br />

dissolving tabular sandstone fragments in the upper<br />

levels <strong>of</strong> the north end <strong>of</strong> the pit, and they<br />

contained virtually no cultural material.<br />

The only feature showing the cultural use<br />

<strong>of</strong> fire consisted <strong>of</strong> a small area about 15 cm in<br />

diameter and about 10 cm deep, running into the<br />

west wall <strong>of</strong> the trench deep in the westward<br />

extension (-8n/-37e). This feature contained<br />

abundant charcoal and ash, but no really associated<br />

artifacts. It is very similar to one encountered in the<br />

1998 excavations <strong>of</strong> Trench 5 (Robb 1998). Its<br />

form suggests it is not natural burning and it may<br />

represent a small fire area used for cooking or for a<br />

special purpose.<br />

The form <strong>of</strong> the site has evolved<br />

substantially during its use. The earliest levels<br />

excavated so far consist <strong>of</strong> Early Neolithic deposits<br />

on top <strong>of</strong> sloping boulders in the north half <strong>of</strong> the<br />

trench; in this period, the site would have afforded<br />

at most about six square meters <strong>of</strong> approximately<br />

level area. Only after enough sediments<br />

accumulated to cover bedrock to the south and<br />

make it more or less level was this area also<br />

occupied. Even so, level area for occupation<br />

consisted <strong>of</strong> at most ca. 15 square meters. This<br />

implies that the site was never used by very large<br />

groups, particularly during the earlier Neolithic,<br />

and that certain functions would have required the<br />

use <strong>of</strong> <strong>of</strong>f-site areas. For instance, the herding <strong>of</strong><br />

animals could have been carried out conveniently in<br />

the basin enclosed by rocky arms <strong>of</strong> the cliffs just<br />

below the site.<br />

Radiocarbon dating and stratigraphic<br />

interpretation. Radiocarbon dates for Trench 1<br />

from the 1998 excavations placed Stratum V in the<br />

6 th millennium BC and Stratum IV in the 5 th<br />

millennium BC, with a date <strong>of</strong> ca. 3000 BC in<br />

Stratum III probably representing an intrusive or<br />

sporadic Copper Age occupation (Table 6).<br />

Following the 1999 season, we obtained two further<br />

dates for Trench 1. One was for the transition<br />

between Strata II and III, and dated it to 4945-4615<br />

BC (Umbro-13, Table 6). The other was for the<br />

center <strong>of</strong> Stratum III, and dated it to 5660-5485 BC<br />

(Umbro-15, Table 6).


These dates are incompatible to some<br />

extent with the chronostratigraphic interpretation<br />

established in 1998, and there are four possible<br />

interpretations.<br />

(1) One is a “high” chronology in which<br />

Strata III-V all date to the 6 th millennium BC,<br />

Stratum II/III date to the 5 th millennium BC. To<br />

argue this, one needs to argue that Diana wares<br />

were in use for most <strong>of</strong> the 5 th millennium BC at the<br />

same time as Stentinello wares, and to discount one<br />

date, Umbro-4, as intrusive into an earlier stratum.<br />

(2) A second is a “low” chronology in<br />

which Stratum V belongs to the 6 th millennium,<br />

Stratum IV belongs to the 5 th millennium, and<br />

Stratum III is a Diana ware level undated but later<br />

than Stentinello times. To argue this, one needs to<br />

discount two dates (Umbro-13 and Umbro-15) as<br />

based upon earlier material remixed into later<br />

strata.<br />

(3) A third is a “middle” chronology in<br />

which Stratum V belongs to the 6 th millennium,<br />

Stratum IV is undated but dates to between the 6 th<br />

and 5 th millennia, and Stratum III dates to the 5 th<br />

millennium. To argue this, one needs to argue that<br />

Stentinello and Diana wares were used together<br />

during the 5 th millennium, and to discount one date,<br />

Umbro-15, as based upon earlier material remixed<br />

into a later level.<br />

(4) The final possibility is that the<br />

stratigraphy is completely mixed and no sense can<br />

be made <strong>of</strong> it.<br />

Of these, the last possibility can be readily<br />

discounted. Sediment pr<strong>of</strong>iles show a distinct<br />

difference at least between Strata II/III and Strata<br />

IV/V, a difference which wholesale mixing would<br />

obliterate. Moreover, Strata IV/V contain no Diana<br />

wares, while Stratum II/III contain both Diana<br />

wares. Again, wholesale mixing <strong>of</strong> levels would<br />

obliterate this difference. The first possibility can<br />

probably also be discounted on archaeological<br />

grounds. Stratum III contains abundant Diana<br />

pottery, and while we might imagine an early Diana<br />

date in the earlier 5 th millennium BC (cf. Leighton<br />

1999), dating a Diana stratum to the earlier 6 th<br />

millennium would place it far earlier than any other<br />

known Diana site (Skeates 1994) and must be an<br />

error.<br />

This leaves the “low” chronology and the<br />

“middle” chronology as possible interpretations.<br />

Both are plausible archaeologically. Given the<br />

shape <strong>of</strong> the site, with Trench 1 intensively<br />

occupied and lying below steep slopes to the south,<br />

it is clear that the stratigraphy was not sealed in<br />

each period but was accumulative. In other words,<br />

early levels contain only early ceramics, charcoal,<br />

and other materials, but later levels contain both<br />

27<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

later materials and earlier material washed in or<br />

mixed in. This provides grounds for understanding<br />

Stentinello pottery mixed in the upper part <strong>of</strong> the<br />

stratigraphy and early dates from upper levels, and<br />

is consistent with both the “low” or “medium”<br />

interpretation.<br />

Here it is worth noting some conclusions<br />

which both views point towards, and some<br />

unresolved issues.<br />

• In both scenarios, Stratum V is a Stentinello<br />

level dating to the 6 th millennium BC, and<br />

Stratum IV is a Stentinello level dating to<br />

sometime between the 6 th and 5 th millennium.<br />

• In both scenarios, Stratum III is a Diana level.<br />

• In both scenarios, Stratum II may be a Copper<br />

Age level with substantial mixing <strong>of</strong> earlier<br />

material, as most <strong>of</strong> the few possible Copper<br />

Age sherds are found in it along with<br />

Stentinello and Diana wares.<br />

• One key unresolved issue is the chronology <strong>of</strong><br />

Stratum IV, which may date to anywhere<br />

between the 6 th and 5 th millennium BC. More<br />

radiocarbon dates from Stratum IV may help<br />

resolve this.<br />

• A second key unresolved issue is the dating <strong>of</strong><br />

the Diana level, Stratum III, and the related<br />

issue <strong>of</strong> whether Diana and Stentinello wares<br />

were in use simultaneously or whether their<br />

association in Stratum III represents mixing <strong>of</strong><br />

earlier and later material. The chronological<br />

issue may be resolvable through further<br />

radiocarbon dates for Stratum III. The ceramic<br />

question may be impossible to resolve with<br />

archaeological evidence from Umbro.<br />

Lest we despair, it should be pointed out that these<br />

problems are normal rather than exceptional.<br />

Virtually every site explored from this period has<br />

similar difficulties. For instance, the regional<br />

cornerstone, the Lipari chronology based upon the<br />

Lipari Acropolis, Contrada Diana, and Castellaro<br />

Vecchio (Bernabò Brea and Cavalier 1956; 1960;<br />

1980), achieved a clear periodization by<br />

discounting numerous sherds in “mixed” levels.<br />

This method assumed in advance <strong>of</strong> the data that<br />

the ceramics seriate into short, clearly-bounded<br />

periods and that any finds to the contrary are<br />

merely contingent errors <strong>of</strong> site preservation which<br />

should be ignored. If one allows other possibilities<br />

such as the contemporary use <strong>of</strong> several styles for<br />

long periods, quite different interpretations are<br />

possible.


3.2.2. Trench 6<br />

Trench 6 proved one <strong>of</strong> the most<br />

interesting surprises <strong>of</strong> the 1999 field season. We<br />

began investigating this area simply to understand<br />

what archaeological deposits existed on the west<br />

side <strong>of</strong> the site. At this point, prehistoric sherds<br />

were visible eroding out <strong>of</strong> a bank about a meter<br />

high, about three meters east <strong>of</strong> the <strong>Bova</strong>-<strong>Marina</strong>-<br />

<strong>Bova</strong> Superior road.<br />

The first effort was to cut a vertical pr<strong>of</strong>ile<br />

along the face <strong>of</strong> this bank; we expected to find<br />

archaeological material tumbled out <strong>of</strong> context at<br />

the base <strong>of</strong> the rocky peak here, but wanted to<br />

check this. A pr<strong>of</strong>ile about three meters long was<br />

cleaned and mapped, running generally north-south<br />

(Figure 11). As this work was completed, it was<br />

clear that most <strong>of</strong> the sherds were localized at one<br />

point where large fragments <strong>of</strong> a single vessel were<br />

visible in the pr<strong>of</strong>ile. We then decided to cut a<br />

trench back from the pr<strong>of</strong>ile to explore further. A<br />

two meter by two meter trench was laid out,<br />

anchored on the site grid. Because the pr<strong>of</strong>ile cut<br />

across the grid at an angle, the actual trench<br />

excavated was a trapezoid including one complete<br />

one meter square and parts <strong>of</strong> two others. Finally,<br />

an extension <strong>of</strong> half a meter thick was opened up on<br />

the eastern and southern margins <strong>of</strong> the trench; this<br />

was excavated only to the upper zones <strong>of</strong> the rock<br />

fall in Stratum II (Figure 12, Figure 15). Trench 6<br />

was excavated by grid squares in 10-cm levels and<br />

all soil was screened in 5 mm mesh.<br />

Five strata were defined in Trench 6<br />

(Figure 13).<br />

Stratum I. This consisted <strong>of</strong> topsoil 5-10<br />

cm thick, a clayey loam (Munsell color 10YR 3/2<br />

dry) with much root penetration and very little<br />

archaeological material.<br />

Stratum II. The basic sediment was the<br />

same as in Stratum I though slightly darker in color<br />

(clayey loam, Munsell color 10YR 5/2 dry). It<br />

contained far less vegetation but much rock fall.<br />

Rocks included both large tumbled stones and<br />

many smaller irregular rocks, tightly packed<br />

together. Overall, the stratum was between thirty<br />

and fifty centimeters thick. <strong>Archaeological</strong> material<br />

was infrequent. This stratum contained two<br />

fragments <strong>of</strong> a ground stone quern (see below),<br />

possibly in secondary context.<br />

Stratum III. This level included a<br />

prehistoric floor or surface. The basic soil<br />

continued to be a sandy, clayey loam (Munsell<br />

10YR 4/2 dry). There was no sign <strong>of</strong> plaster or<br />

paving, but the soil was compacter than above;<br />

there were fewer large rocks but a dense<br />

28<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

concentration <strong>of</strong> small stones; many small stones<br />

were lying flat; and most <strong>of</strong> the few sherds found<br />

scattered through the level were lying flat. This<br />

level was about fifteen centimeters thick. It was<br />

excavated away only in the westernmost part <strong>of</strong> the<br />

Trench, and left in situ for further excavation in the<br />

eastern areas. The radiocarbon samples for Trench<br />

6 came from this stratum.<br />

Stratum IV. This stratum was visible only<br />

at the western edge <strong>of</strong> the pit where erosion and<br />

excavation had exposed it. It appears to be a<br />

whitish, sterile soil <strong>of</strong> variable thickness (a greyishwhite<br />

concreted sand and clay, Munsell color 10YR<br />

5/2). It appears to contain a few pieces <strong>of</strong><br />

archaeological material but may be the top <strong>of</strong> sterile<br />

sediments derived from decomposing bedrock.<br />

Stratum V. Bedrock.<br />

Pot deposition and other features. As we<br />

excavated Trench 6, the vessel appearing in the<br />

pr<strong>of</strong>ile turned out to be part <strong>of</strong> a group <strong>of</strong> three pots<br />

(Figure 14, Figure 16). The pot visible in the<br />

pr<strong>of</strong>ile was a basin between thirty and forty<br />

centimeters in maximum diameter, apparently<br />

undecorated on the surface but with three<br />

horizontal handles under the rim. Its overall form is<br />

yet to be understood as it was extensively crushed.<br />

The other two pots consisted <strong>of</strong> small attingitoi or<br />

cup-dippers with tall upraised handles. One was<br />

relatively thin-walled, the other cruder and thicker.<br />

The three pots were placed close together, with<br />

their walls almost touching. Their bases rested on<br />

the same plane at 70 cm below trench datum, a<br />

place corresponding with the floor level defined in<br />

Stratum III. As these pots were excavated, they<br />

were mapped and photographed, and their fill was<br />

collected completely as a series <strong>of</strong> flotation samples<br />

(see below). Under the two attingitoi, we found a<br />

fragment <strong>of</strong> a fourth pot, which consisted only <strong>of</strong><br />

one horn <strong>of</strong> a horned bowl.<br />

Large and small rock fall covered the<br />

entire area including the pot deposition group, but<br />

grew thicker to the east. Larger stones were more<br />

common in the southeastern corner <strong>of</strong> the square.<br />

No order was visible in the rock fall. No other<br />

possible features or structures were visible, though<br />

the bedrock at the southwestern corner <strong>of</strong> the trench<br />

may have been artificially shaped to some extent.<br />

Dating. Both the vessel styles and absolute<br />

datations agree in assigning the deposition to the<br />

Early Bronze Age. While analysis <strong>of</strong> the large bowl<br />

must await its reconstruction, the form <strong>of</strong> the<br />

attingitoi is very similar to ones from Mursia,<br />

Pantelleria, varied Rodi-Tindari-Vallelunga sites<br />

such as Ciavolaro in Sicily, and S. Domenico<br />

Ricadi in Calabria. The ceramic horn is distinctive<br />

<strong>of</strong> the Rodì-Tindari-Vallelunga style (Tusa 1993;


Castellana 1996). A radiocarbon sample from<br />

charcoal on the surface below the vessel group gave<br />

a date <strong>of</strong> 3390 +/- 60 BP (1780-1520 BC<br />

calibrated).<br />

Because <strong>of</strong> the limited excavations carried<br />

out at this location to date, it is not entirely clear<br />

what kind <strong>of</strong> deposition this group <strong>of</strong> pots<br />

represents. They may represent the remnant <strong>of</strong> a<br />

house context. The daub fragments associated with<br />

them may argue in favor <strong>of</strong> this, but the finds from<br />

the trench are far sparser than one would expect<br />

from a domestic context, including virtually no<br />

animal bone or lithics, and there is yet no sign <strong>of</strong><br />

any structure such as the typically substantial stonefooted<br />

Bronze Age huts known from eastern Sicily<br />

and Lipari. A second possibility is ritual: the pots<br />

may represent an idiosyncratic ritual deposition <strong>of</strong><br />

some kind, though no comparable sites have been<br />

published. At this point, burial seems an equally<br />

plausible possibility. In eastern Sicily, EBA burials<br />

are commonly found in cemeteries <strong>of</strong> small rockcut<br />

tombs, located along elevated ridges or cliffs <strong>of</strong><br />

limestone or other calcareous stone. Umbro is in<br />

just such a location, and the vessel group lies 2-3<br />

meters from an outcropping, low bedrock ridge,<br />

now buried. Furthermore, in several cases, outside<br />

rock-cut tombs archaeologists excavated leveled<br />

areas or platforms which contained groups <strong>of</strong><br />

pottery vessels, interpreted as grave <strong>of</strong>ferings<br />

deposited outside tombs rather than inside (see for<br />

instance Santa Febronia, Maniscalco 1996). Pottery<br />

vessels used as grave goods in the Bronze Age<br />

<strong>of</strong>ten consisted <strong>of</strong> groups including a bowl or basin<br />

and several cups or dippers, interpreted as a<br />

complete pottery service used for ritual functions<br />

(Maniscalco 1999). This is precisely the<br />

composition <strong>of</strong> the Umbro pot group. It is possible<br />

that there are rock-cut tombs in the rocky bank just<br />

behind the pottery group, and that the pots<br />

represent external grave goods for such tombs, with<br />

the rock fall representing remnants <strong>of</strong> a wall closing<br />

the tombs (B. McConnell and L. Maniscalco, pers.<br />

comm.). Needless to say, we hope to excavate<br />

further in this area, to ascertain the actual nature <strong>of</strong><br />

the site.<br />

3.2.3. Trench 7<br />

Trench 7 continues the exploration <strong>of</strong><br />

outlying parts <strong>of</strong> the site already begun with Area 2,<br />

Trench 3, Trench 4 and Trench 5.<br />

The trench consisted <strong>of</strong> an exploratory one<br />

meter square pit (Figure 17, Figure 19). It was<br />

located on a small (ca. 4 meters in either direction)<br />

level area at the very crest <strong>of</strong> the southern slope<br />

below the cliffs, just below the small field<br />

described below and next to the “south slope<br />

29<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

pr<strong>of</strong>ile” (see Robb 1998). The “cliffs” above it<br />

were only about two meters tall, consisting <strong>of</strong> very<br />

irregularly sloping bedrock outcrops. It was located<br />

about six meters southeast <strong>of</strong> Trench 3. This level<br />

patch, one <strong>of</strong> very few on the hillside, had been<br />

used in the previous year for a backdirt heap for<br />

Trench 3, and some soil remaining after Trench 3<br />

was backfilled was removed without screening. By<br />

these means, the terrace was cleared to a uniform<br />

surface, except for a low, flat boulder in its center.<br />

Beyond a few historic sherds clearly in secondary<br />

context, no artifacts were evident either in this soil<br />

removal or in the surface thus cleared. Once the<br />

working ground was cleared, before laying out the<br />

trench, we cleaned away some lose, irregular soil<br />

adhering to the cliff wall directly above where we<br />

wished to place the trench; this was to prevent this<br />

soil from falling into the trench later. About half a<br />

cubic meter <strong>of</strong> soil was removed in this way,<br />

leaving a vertical surface. Again, only a few<br />

historic sherds out <strong>of</strong> context were found.<br />

Trench 7 was located directly against the<br />

rocky wall rather than in the center <strong>of</strong> the level,<br />

open area. Trench 3 had shown that such level<br />

pockets adhering to steep slopes could contain<br />

considerable depth <strong>of</strong> soil, but that they were <strong>of</strong>ten<br />

stratigraphically complex. Moreover, there were no<br />

indications on the surface <strong>of</strong> the level area <strong>of</strong><br />

underground deposits here. A primary<br />

consideration, however, was that <strong>of</strong> caves. The<br />

shelving, angled bedrock at this point creates the<br />

possibility <strong>of</strong> small pocket caves in the cliff wall.<br />

With increasing evidence that the cliff-tops<br />

immediately above were not settled, and that the<br />

hillslope deposits such as in Trench 3 and Area 2<br />

are not in original context, we wondered whether<br />

prehistoric hillside deposits might originally have<br />

been in small caves or shelters, now buried or<br />

collapsed, high along the rocky walls <strong>of</strong> the basin.<br />

Trench 7 was excavated using the same<br />

methods as elsewhere: excavation with trowel,<br />

handpick, and pick proceeded in 10 cm arbitrary<br />

levels, and all soil was screened in 5 mm mesh. A<br />

trench datum was set up using the cleft in the large<br />

flat boulder north <strong>of</strong> the trench to lodge a datum<br />

spike; this was the only secure point available from<br />

which depths could be practically measured.<br />

Excavation continued to a total depth <strong>of</strong> 150 cm<br />

below trench datum, or ca. 110 cm below the<br />

cleared ground surface.<br />

Two strata were defined in the pit, both <strong>of</strong><br />

extremely irregular dimensions (Figure 18). The<br />

upper one was a very loose loam containing sand,<br />

silt and clay. It contained stones <strong>of</strong> all dimensions<br />

jumbled without consistent orientation, and was<br />

much disturbed by rootlets. The maximum depth<br />

was about 120 cm below trench datum, or close to


the bottom <strong>of</strong> the pit. This sediment clearly<br />

represents soil filtering in among the rocks <strong>of</strong> the<br />

cliffs, a process visible everywhere on the cliff<br />

slopes. It did not appear internally stratified in any<br />

sense, and it is not clear when it was deposited or<br />

formed. It contained a few sherds from all dates<br />

including prehistoric, medieval and modern. The<br />

second, lower stratum was very similar in its basic<br />

substance, but was slightly less loose in texture,<br />

slightly browner. Again, it contained both<br />

prehistoric and historic sherds. It thus seems to be<br />

an ancient version <strong>of</strong> soil and rocky infill <strong>of</strong><br />

irregularities in the cliffs.<br />

Trench 7 finds included only ceramics.<br />

The density <strong>of</strong> finds was very low and three levels<br />

produced no artifacts at all. Prehistoric and historic<br />

sherds were mixed down to the bottom <strong>of</strong> the<br />

trench, suggesting that all the fill excavated was<br />

mixed and <strong>of</strong> recent date.<br />

We stopped excavating Trench 7 because<br />

the field season ended, and because it was not clear<br />

how to continue excavating safely in a small trench<br />

more than a meter deep in loose deposits partially<br />

below a rocky overhang. There was no indication<br />

how deep soil deposits ultimately run here on the<br />

cliff-side; if we assume that the cliffs were<br />

originally vertical all around the sides <strong>of</strong> the basin<br />

enclosing the Neolithic site, and that the current<br />

slopes <strong>of</strong> approximately forty-five degrees have<br />

been formed by millennia <strong>of</strong> rock fall and soil infill,<br />

there could potentially be twenty meters <strong>of</strong> deposits<br />

here.<br />

With this in mind, we need to return to the<br />

question <strong>of</strong> how to check for possible cave deposits<br />

in the rocky walls <strong>of</strong> the basin. Figuring out how to<br />

investigate this safely and practically remains a<br />

challenge for the future.<br />

3.3. Other Areas Explored<br />

With the discovery <strong>of</strong> distinct<br />

archaeological areas on the west side <strong>of</strong> the site<br />

above the cliffs, and with continued questioning<br />

about the nature <strong>of</strong> the upper stratigraphy <strong>of</strong> Trench<br />

1 below the cliffs, we decided it was important to<br />

re-check whether there were any archaeological<br />

deposits in the intervening parts <strong>of</strong> the site – the<br />

highest peak <strong>of</strong> the rocky cliffs. In particular, we<br />

were concerned to know whether there was any<br />

possibility that the artifacts in Trench 1 and<br />

elsewhere below the cliffs had fallen there from<br />

cliff-top deposits, and whether the newly-found<br />

Bronze Age deposits in Trench 6 were in situ, as<br />

they seemed to be, or could have eroded from<br />

deposits higher on the peak.<br />

30<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

We had, <strong>of</strong> course, thoroughly examined<br />

the entire ground surface <strong>of</strong> these peaks in previous<br />

years, and we did so again. Perhaps half the surface<br />

consists <strong>of</strong> irregular, shelving bedrock outcrops;<br />

between these are pockets <strong>of</strong> soil <strong>of</strong> varying depths,<br />

and the whole area is sparsely covered in scrubby<br />

vegetation. Ground visibility was thus patchy but<br />

reasonable in places. All previous examinations had<br />

failed to find any archaeological material <strong>of</strong> any<br />

date.<br />

This season, we decided to check for<br />

possible archaeological deposits below the surface<br />

(for instance, surviving in bedrock crevices). Since<br />

there were no known archaeological deposits in this<br />

area, and since the soil covering the area consisted<br />

<strong>of</strong> thin, irregular pockets, we excavated it<br />

informally by cleaning the bedrock <strong>of</strong>f with a<br />

shovel. A team <strong>of</strong> L. Saccà and a student carried<br />

out the work. Most <strong>of</strong> the soil removed was<br />

screened, but not all was, as it rapidly became<br />

apparent that the sediments were entirely sterile and<br />

that the team could monitor it adequately in the<br />

process <strong>of</strong> removing it. This strategy carried the<br />

risk <strong>of</strong> damaging real archaeological deposits<br />

through loosely controlled excavation, but we felt<br />

that any deposits encountered could be recognized<br />

rapidly enough to stop, grid, and excavate in a<br />

controlled way with only limited damage;<br />

meanwhile, we needed rapid information without<br />

committing ourselves for a full-scale excavation in<br />

an area <strong>of</strong> patchy, sterile soil and bedrock. All<br />

cleaned areas were backfilled when finished. In the<br />

event, this strategy was justified.<br />

We cleaned the bedrock in five areas,<br />

three on the high area <strong>of</strong> the cliff-top, one along the<br />

cliff edge above Trench 3, and one to the southeast<br />

above Trench 7. In all <strong>of</strong> them, the soil consisted <strong>of</strong><br />

a loose, dark loam with much organic material.<br />

Atop the cliffs, the northernmost area<br />

consisted <strong>of</strong> the largest pocket <strong>of</strong> soil close to the<br />

actual cliff edge which could be safely worked at. It<br />

lay directly above Trench 1, about three meters in<br />

from the actual lip <strong>of</strong> the cliff. The soil filled a<br />

shallow trough about two meters east-west and<br />

perhaps .75 meters north-south. The deepest part <strong>of</strong><br />

the excavation was about 40 cm deep. The other<br />

two areas cleared atop the cliffs were located<br />

southwest <strong>of</strong> here, on the very highest point <strong>of</strong> the<br />

hill. These consisted <strong>of</strong> two parallel troughs about<br />

four meters long and less than a meter wide, each<br />

filling in a v-shaped recess created by shelving<br />

bedrock (Figure 20). The maximum soil depth in<br />

both was 30 cm. None <strong>of</strong> these trenches contained<br />

any artifacts.<br />

Along the cliff edge above Trench 3, we<br />

cleared a thin covering <strong>of</strong> loose topsoil from


edrock in a series <strong>of</strong> small pockets. None<br />

contained any cultural material.<br />

Southeast <strong>of</strong> the rocky peak, there is a<br />

small level field which lies above the southern<br />

slope <strong>of</strong> the basin between the cliffs. The pr<strong>of</strong>ile<br />

just below this field was cleaned and recorded in<br />

1998 (“South slope pr<strong>of</strong>ile”, Robb 1998), and<br />

Trench 7 was excavated just below it in 1999. If<br />

archaeological material was washing into the basin<br />

from somewhere high on the southern slopes, this<br />

seemed one candidate for a source area. We did not<br />

clean this area down to bedrock, as the soil was <strong>of</strong><br />

unknown depth and we were reluctant to intervene<br />

here substantially without setting up a formal<br />

excavation area. However, we cleared the ground<br />

vegetation and removed the top 10 cm <strong>of</strong> soil using<br />

the same techniques as above. The only finds<br />

included a number <strong>of</strong> sherds from a single modern<br />

vessel. This confirms the generally sterile nature <strong>of</strong><br />

this part <strong>of</strong> the site inferred from the south slope<br />

pr<strong>of</strong>ile and from Trench 7.<br />

These exercises were very useful in<br />

understanding the site. It seems virtually certain<br />

that the Trench 1 deposits and the south slope<br />

deposits in the area <strong>of</strong> Trench 3 were probably<br />

originally deposited below the cliffs; it is difficult<br />

to imagine substantial enough deposits above the<br />

cliffs which would not leave at least some residual<br />

archaeological material up above. Likewise, the<br />

Bronze Age deposits in Trench 6 seem localized<br />

and in their original location. The presence <strong>of</strong> an<br />

elevated, archaeologically empty area between the<br />

Neolithic and Bronze Age areas confirms the<br />

separation <strong>of</strong> these deposits.<br />

While on the topic <strong>of</strong> other areas<br />

investigated, it is worth illlustrating some <strong>of</strong> the<br />

1998 trenches briefly; these trenches were<br />

described in the previous preliminary report (Robb<br />

1998) but no photographic docmentation was then<br />

available. Figure 21 shows Area 2, a zone on the<br />

southern slope <strong>of</strong> the basin where we cleaned a<br />

pr<strong>of</strong>ile within a cleft between tumbled boulders but<br />

found only redeposited material in secondary<br />

position. Figure 22 shows Trench 3, a one meter by<br />

one meter test pit further up the same slope, and<br />

Figure 24 depicts Trench 5, a three square meter<br />

trench located below Trench 1 in the very bottom<br />

<strong>of</strong> the basin. The boulders looming over the trench<br />

are those underlying Trench 1. Finally, Figure 23<br />

depicts Trench 4, a broad, shallow four by five<br />

meter pit located on a rocky outcrop above the<br />

cliffs. It yielded no structures or architecture, but a<br />

small assemblage <strong>of</strong> post-Neolithic pottery which is<br />

probably Bronze Age in date.<br />

31<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


4. UMBRO EXCAVATIONS: DESCRIPTION OF FINDS<br />

In this section, finds from the Umbro<br />

excavations are discussed. Note that this discussion<br />

includes both new finds from the 1999 excavations<br />

and previously described finds from the 1998<br />

excavations; the reason is that our understanding <strong>of</strong><br />

earlier finds has <strong>of</strong>ten changed as research has<br />

progressed.<br />

4.1. Ceramics<br />

4.1.1. Neolithic and Copper Age wares: Trench 1.<br />

The typical wares found in the Neolithic<br />

area <strong>of</strong> Umbro have already been described (Robb<br />

1998), and pending statistical analysis, little<br />

additional comment can be made. They will only be<br />

briefly presented here. Pottery styles found in the<br />

area below the cliffs at Umbro include Stentinello<br />

coarse wares and finewares, Diana wares, and<br />

possible Copper Age sherds.<br />

Stentinello coarse wares (Figure 25) are<br />

<strong>of</strong>ten thick-walled and relatively crudely made,<br />

executed in dark brown, light brown, red and<br />

orange fabrics. Some vessels in this category may<br />

have been bowls or cooking pots, while others<br />

appear to have been large storage vessels. They<br />

were decorated with a range <strong>of</strong> impressed designs,<br />

including “stab and drag” punctuation, shell and<br />

rocker impressions, and arrays <strong>of</strong> short parallel<br />

lines, <strong>of</strong>ten arranged vertically. They are similar not<br />

only to coarse wares at other Stentinello sites such<br />

as Acconia and Capo Alfiere, but to coarser<br />

Impressed Ware vessels in general.<br />

Stentinello finewares (Figure 25) are<br />

usually thinner-walled, with dark, glossy burnished<br />

surfaces apparently intended to highlight the<br />

impressions filled with yellow, red or white pastes.<br />

Decoration was by impression, with geometrical<br />

arrays <strong>of</strong> small motifs forming elaborate design<br />

schemes. The most common motifs were “v”<br />

impressions, small straight lines and fine grids <strong>of</strong><br />

diagonal lines. These were recombined in bands<br />

below the rim <strong>of</strong> a vessel, with vertically “hanging”<br />

strips <strong>of</strong> decoration or further banding below.<br />

Decoration also occurred elsewhere on vessels, as<br />

shown by a small handle with an elaborate strip <strong>of</strong><br />

impressions down its back. Many <strong>of</strong> these vessels<br />

appear to have been small open bowls, probably for<br />

serving food and drink.<br />

Diana wares (Figure 26) include both<br />

coarse and fine fabrics, though the coarse fabrics<br />

can be difficult to distinguish from other Neolithic<br />

32<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

wares. The Umbro Diana finewares include many<br />

shallow carinated bowls with reddish, brownish or<br />

buff fabrics, highly burnished, occasionally slipped<br />

with a red slip, and usually undecorated save for a<br />

thin line below the rim and for the famous trumpet<br />

handles attached horizontally below the rim.<br />

In addition to Diana and Stentinello wares,<br />

other Neolithic wares included a few red-painted<br />

sherds on a buff fabric, including some from<br />

relatively deep in Stratum IV.<br />

A few sherds unusual for Neolithic wares<br />

(though not entirely unparalleled) and with generic<br />

Copper Age parallels were found in the 1998<br />

excavations (Figure 27). These included several<br />

body sherds with large circular bosses, and several<br />

large horizontal handles lacking the typical Diana<br />

flaring. More such sherds were found in 1999, and<br />

the greater extent <strong>of</strong> excavation allowed us to<br />

define their stratigraphic level as Stratum II. These<br />

included several large, thick body sherds with a<br />

double band <strong>of</strong> finger-pinched cordoned<br />

decoration, perhaps from the same vessel. Another<br />

sherd was incised with a series <strong>of</strong> evenly-spaced<br />

vertical lines ending in spirals at the top. An<br />

especially unusual fragment was decorated with a<br />

broad, shallow groove interrupted by a pair <strong>of</strong><br />

protuberances. Finally, several sherds bore incised<br />

designs consisting <strong>of</strong> two parallel zig-zag lines, the<br />

space between which was filled with rough<br />

hatching (cf. Cocchi Genick 1996 for parallels from<br />

the Southern Italian Copper Age).<br />

4.1.2. Stampini and other ceramic-production<br />

related implements<br />

The Umbro excavations have yielded six<br />

stampini, or small ceramic punches, to date (Figure<br />

29). They consist <strong>of</strong> small fired clay cylinders with<br />

simple geometrical motifs in raised clay at the end<br />

or on the side. These small tools were used to<br />

impress unfired vessels to create the stereotypical<br />

Stentinello pottery designs. At Umbro, they have<br />

been found in all strata <strong>of</strong> Trench 1. Examples<br />

include:<br />

• a complete stampino with a v-motif at one end;<br />

the edges <strong>of</strong> the v are finely ticked (Bag 1021;<br />

length 70 mm).<br />

• a fragmentary handle <strong>of</strong> a stampino (Bag<br />

1035).


• a fragmentary stampino, approximately 48 mm<br />

long and 10 mm in diameter and slightly<br />

broken; the end was flared but the exact motif<br />

was eroded and indistinct (Bag 1152).<br />

• a fragmentary stampino, whose end had broken<br />

<strong>of</strong>f. The surviving end appears to have been<br />

triangular, suggesting that the motif on the end<br />

was originally a “v” (Bag 1170).<br />

• a fragmentary stampino ending in an eroded<br />

but distinct “v” impression (Bag 1195).<br />

• a fragmentary stampino with a cylindrical<br />

body. Unusually, this fragment bore a raised<br />

motif for impressing clay on the side <strong>of</strong> the<br />

body rather than the end. The motif consisted<br />

<strong>of</strong> a thin ridge <strong>of</strong> clay forming a straight line<br />

with tick marks along its edge (Bag 1225).<br />

These numerous tools for making pots suggest that<br />

ceramic production was a common activity at<br />

Umbro in the Stentinello period. This impression is<br />

reinforced by the presence <strong>of</strong> other implements<br />

which could have been used for forming or<br />

decorating pots, including a shark’s tooth, a<br />

smoothed bone, several small marine shells with<br />

worked edges. Finally, one large, reddish sherd had<br />

two edges worn round rounded and smoothed by<br />

protracted use in smoothing or polishing some<br />

material (Figure 29). All <strong>of</strong> these could have been<br />

used in forming, decorating or burnishing pots prior<br />

to firing them.<br />

4.1.3. Thin section analysis <strong>of</strong> prehistoric ceramics<br />

Fifteen sherds were transported to the<br />

University <strong>of</strong> Southampton for thin section analysis<br />

(we are grateful to the Soprintendenza<br />

Archeological della Calabria for permission to<br />

carry out this work). These sherds represented a<br />

variety <strong>of</strong> prehistoric wares in use at the site,<br />

including Stentinello coarse wares (sample 4) and<br />

decorated (2, 3, 9, 11) and undecorated (7, 8, 15)<br />

fine wares, Diana vessels (10, 13, 14), Early<br />

Bronze Age vessels, possible Copper Age vessels<br />

(6, 12), and red-painted wares (1) which were<br />

considered possible imports. A piece <strong>of</strong> daub was<br />

also examined (5). Finally, samples <strong>of</strong> clay from<br />

two clay sources were examined. One clay sample<br />

was from the field directly east <strong>of</strong> the site at a<br />

distance <strong>of</strong> 10 meters east <strong>of</strong> the site datum point.<br />

The other was taken from a pure clay bank in the<br />

<strong>Bova</strong>-<strong>Bova</strong> Superiore road cut just south <strong>of</strong> the<br />

Umbro plateau where the road begins to rise<br />

steeply, about 500 m south <strong>of</strong> the Neolithic site.<br />

The thin-section analysis was carried out<br />

by Sonya Collins and Jayne Watts under the<br />

supervision <strong>of</strong> Dr. David Williams; we are grateful<br />

to these researchers for their work, discussed here<br />

33<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

and reproduced in the attached documentation. All<br />

sherds were impregnated with a consolidating glue<br />

and then thin-sectioned. Sherd 1 proved too s<strong>of</strong>t<br />

and crumbly to thin-section, but was<br />

macroscopically described and illustrated. The clay<br />

samples were worked into lumps, fired,<br />

impregnated with consolidating glue and thinsectioned.<br />

Thin section analyses, while preliminary,<br />

suggest a number <strong>of</strong> trends.<br />

First, virtually all the sherds had the same<br />

composition. Common mineral inclusions in the<br />

pastes include mica, quartz, and feldspar, with<br />

possible metamorphic rock sources including gneiss<br />

and micaschist. All <strong>of</strong> these are consistent with the<br />

kinds <strong>of</strong> minerals to be found locally resulting from<br />

the decomposition <strong>of</strong> metamorphic rocks common<br />

in Aspromonte at altitudes <strong>of</strong> 500 meters and<br />

above. Hence, these minerals are available, for<br />

instance, in the form <strong>of</strong> sand in river valleys, and by<br />

crushing up parent rock from nearby outcrops. The<br />

only exceptions were the daub sample (5) and a rim<br />

(14). The daub contained similar mineral inclusions<br />

but included more <strong>of</strong> them and larger grains.<br />

Sample 14, a small, fine greyish Diana rim with the<br />

characteristic tubular handle, contained an unusual<br />

base clay with abundant foraminifera fossils and<br />

contained mica, quartz, and black iron grains.<br />

Given the high mica content, it seems likely that it,<br />

too, came from some area <strong>of</strong> Aspromonte, but it<br />

may have been made in a different area, from<br />

different raw clay, or using different temper.<br />

The fact that all the other sherds had a<br />

common composition suggests that they were all<br />

locally made, even when they are made in very<br />

different styles such as red-painted wares,<br />

originally thought to be a potential import from<br />

Northern Calabria. This emphasis on local<br />

production is a common finding in prehistoric<br />

Italian thin-section studies (for instance, Skeates<br />

1992).<br />

The second surprising finding emerged<br />

from comparison <strong>of</strong> the pottery with clay sources<br />

near the site. Both clay sources near the site<br />

contained high levels <strong>of</strong> organic elements in the<br />

form <strong>of</strong> micr<strong>of</strong>ossils. They were clearly<br />

distinguishable from the clays used in all the sherds<br />

except for sample 14 (discussed above) and to a<br />

lesser extent sample 15. This suggests that,<br />

although there is strong evidence that people were<br />

making pottery at Umbro (see below), and although<br />

large quantities <strong>of</strong> apparently suitable clay are<br />

available within a few meters <strong>of</strong> the site, people<br />

preferred to transport clay from somewhere else to<br />

make pots. This is especially interesting as regards<br />

the daub sample, as one would expect a bulky<br />

material such as daub to have been made from the


closest available clay to save work. It may be that<br />

the potters <strong>of</strong> Umbro knew the qualities <strong>of</strong> the local<br />

clays intimately and preferred to choose the most<br />

suitable for a particular task even if they were not<br />

the closest to hand.<br />

These investigations require further<br />

samples and further contextualization in the local<br />

geology and should be considered preliminary<br />

indications only.<br />

4.1.4. Bronze Age wares: Trench 6.<br />

Ceramics from Trench 6 were completely<br />

different from the Neolithic wares found elsewhere<br />

on the site. The general paste was s<strong>of</strong>t, coarse and<br />

light, ranging from orange to brown in color and<br />

usually with a black oxidation stripe in the center <strong>of</strong><br />

the fabric. There was no surface decoration evident<br />

beyond an occasional groove or ticking along a rim.<br />

The repertory <strong>of</strong> vessel forms was also distinct,<br />

with everted rims, thick, beaded lips, ring bases,<br />

and tall raised strap handles. These wares closely<br />

resembled pottery found in Trench 4, and the two<br />

probably date to the same period, the Early Bronze<br />

Age.<br />

In addition to these wares, a few sherds <strong>of</strong><br />

a very different style were found, <strong>of</strong> a black fabric<br />

with a very glossy highly burnished surface. These<br />

sherds represented at least two vessels. Two sherds<br />

were decorated with broad, shallow grooves. These<br />

wares may be similar to Apennine pottery, although<br />

the grooves are not excised. Alternatively, they may<br />

be the wares which Tinè (1992) considered similar<br />

to Maltese Borg-in-Nadur wares.<br />

The pottery deposition from Trench 6<br />

contained three vessels and a fragment <strong>of</strong> a fourth.<br />

These vessels are currently under restoration by<br />

conservators <strong>of</strong> the Museo Nazionale di Reggio<br />

Calabria.<br />

• Vessel 1 (Figure 28) was a small, thin-walled<br />

cup-dipper (attingitoio). The fabric was a s<strong>of</strong>t<br />

buff/orange inside and outside, with fine grit<br />

temper and a thin oxidation stripe. There was<br />

no surface decoration. The lip was everted, and<br />

a tall handle was raised directly from the lip.<br />

Both this cup and Vessel 2 resemble in form,<br />

though not necessarily in fabric or surface<br />

treatment, contemporary vessels from Mursia<br />

(Pantelleria), Capo Graziano, Rodi-Tindari-<br />

Vallelunga sites, and S. Domenico Ricadi<br />

(Tropea)(Bernabò Brea 1957; Castellana 1996;<br />

Tusa 1993).<br />

• Vessel 2 (Figure 28) was a small, thick-walled<br />

cup-dipper (attingitoio). The fabric was dark<br />

brown externally and orange-brown internally,<br />

with a black oxidation stripe and small grit<br />

34<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

inclusions. The surface showed occasional<br />

signs <strong>of</strong> having been burnished, but is now<br />

eroded. There was a shallow groove the<br />

outside surface under the rim.<br />

• Vessel 3 was a large impasto bowl, executed in<br />

a s<strong>of</strong>t orange-brown fabric. There was no<br />

observable surface decoration, though the<br />

surface was eroded and in places covered with<br />

a tenacious concretion. The precise form <strong>of</strong> the<br />

vessel is not yet known, as it is still under<br />

reconstruction, but it measured 20-25 cm in<br />

diameter at the mouth, 10-15 cm in diameter at<br />

the base, and was 15-20 cm tall. Below the rim<br />

there were three horizontal strap handles.<br />

• Below these was found a fragment <strong>of</strong> a fourth<br />

vessel, a ceramic horn. This was made <strong>of</strong> a<br />

dark brown, burnished fabric. The concave<br />

side was flattish, while the convex side <strong>of</strong> the<br />

horn was rounded. This cross-section makes it<br />

resemble more closely the horns from the<br />

Rodi-Tindari-Vallelunga style horned cups<br />

(see Castellana 1996) than the later Ausonian<br />

wares.<br />

4.2. Lithics<br />

Lithics in 1999 were found only in Trench<br />

1. One flake was found in Trench 6, but it was a<br />

crude quartzite flake and may be natural rather than<br />

humanly manufactured. The large number <strong>of</strong> lithics<br />

found in Trench 1 brings the total lithic assemblage<br />

there to 401 items (Table 3). Of these, 17 (4.2%)<br />

were made from flint; 384 (95.8%) were made from<br />

obsidian. This overwhelming use <strong>of</strong> obsidian is<br />

typical <strong>of</strong> Stentinello sites in southern Calabria,<br />

eastern Sicily and Lipari (see Ammerman 1987),<br />

though obsidian use falls to about 50% in the<br />

further margins <strong>of</strong> Stentinello culture area (e.g.<br />

Stentinello, Orsi 1890; Capo Alfiere, Morter 1992).<br />

To date, only a small sample <strong>of</strong> about<br />

twenty pieces has been examined by a lithics<br />

analyst (G. Marshall, see Robb 1998); the finds<br />

from the 1999 season do not appear to change the<br />

conclusions <strong>of</strong> this analysis. A full analysis will<br />

take place during the 2001 study season. Marshall’s<br />

analysis showed that obsidian probably entered<br />

Umbro in the form <strong>of</strong> small blade cores, which<br />

were then used to produce blades until they were<br />

too small, at which point they were were crushed<br />

using bipolar percussion to create small flakes for<br />

expedient use.<br />

Obsidian provenience analysis in progress<br />

at the University <strong>of</strong> South Florida has shown that<br />

all samples examined from Umbro came from<br />

Lipari (R. Tykot, pers. comm.). Flint came from a<br />

variety <strong>of</strong> sources, as at least five distinct colors <strong>of</strong>


aw material are known (grey, brown, yellow, red,<br />

and pink). While grey flint <strong>of</strong> poor quality is found<br />

at Saracena in the hinterlands <strong>of</strong> Cond<strong>of</strong>uri <strong>Marina</strong>,<br />

the sources <strong>of</strong> the other varieties are unknown. As<br />

elsewhere in Calabria, Sicily and Lipari, obsidian<br />

was used primarily for bladelets and unmodified<br />

flakes; flint was used disproportionately for formal<br />

tools and modified flakes, probably for tasks which<br />

required its somewhat tougher cutting edge.<br />

4.3. Other Artifacts<br />

4.3.1. Daub<br />

Daub, or intonaco, is a common Neolithic<br />

technology. To date, 1172 daub fragments have<br />

been recovered at Umbro (Table 4). Of these, 3<br />

came from Trench 3, 5 from Trench 4, 4 from<br />

Trench 5, and 46 from Trench 6. These serve to<br />

document the use <strong>of</strong> daub at least in the Trench 6<br />

area. The vast majority <strong>of</strong> fragments (1142) come<br />

from Trench 1, the Neolithic area <strong>of</strong> the site.<br />

Almost all <strong>of</strong> the daub fragments<br />

recovered consist <strong>of</strong> tiny fragments less than a<br />

centimeter or so in maximum dimension. Only<br />

about 80 pieces are larger. The largest fragment<br />

found is about 10 cm. No daub fragments were<br />

found in original position; all were scattered<br />

through the soil, though there were slightly greater<br />

concentrations between 150-180 cm below datum<br />

(in the lower part <strong>of</strong> Stratum II and in Stratum III)<br />

in the central part <strong>of</strong> the trench.<br />

All fragments found were burnt to some<br />

degree, usually only lightly but intensely in a few<br />

cases. This does not necessarily imply that burning<br />

was part <strong>of</strong> the use <strong>of</strong> daub here; rather, burning<br />

probably helped preserve daub and daub fragments<br />

which were not burned at some point simply<br />

disintegrated and were not preserved. However,<br />

some burning clearly happened before the daub was<br />

completely fragmented and quite possibly during its<br />

use-life; this is suggested by a number <strong>of</strong> flat<br />

fragments consistently burned on one side only (see<br />

below).<br />

A handful <strong>of</strong> fragments display<br />

impressions <strong>of</strong> sticks or reeds; the largest<br />

impression is <strong>of</strong> a stick about four centimeters in<br />

diameter, but impressions between one and two<br />

centimeters are more common.<br />

Between twenty and thirty fragments<br />

appear to come from a single artifact or structure.<br />

These fragments are flat Running cross-wise<br />

through these fragments in places are impressions<br />

<strong>of</strong> three reeds or sticks about one centimeter in<br />

diameter, laid edge to edge. The daub thus formed a<br />

flattish sheet about five centimeters thick, enclosing<br />

on a framework <strong>of</strong> three parallel reeds with about a<br />

35<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

centimeter to spare on the inside and outside. One<br />

side was smoothed, and occasionally betrays<br />

possible plank impressions or smoothing marks.<br />

The other side is roughened. The rough side was<br />

consistently burned to a blackish color; the<br />

smoothed side was burned far less intensely, and at<br />

most shows some reddening.<br />

Daub is stereotypically interpreted as<br />

having been used for house construction. However,<br />

sites <strong>of</strong> Neolithic houses built in daub <strong>of</strong>ten yield<br />

very large amounts <strong>of</strong> it (Shaffer 1985; Tozzi and<br />

Tasca 1989). There is no signs <strong>of</strong> such large<br />

quantities <strong>of</strong> daub at Umbro, nor is it clear that<br />

there would have been room for a house even 3-4<br />

meters square in the constricted area below the<br />

cliffs at Trench 1. Rather, the daub may have been<br />

used in some other kind <strong>of</strong> artifact or facility. The<br />

group <strong>of</strong> flat daub fragments may belong to some<br />

feature such as a partition, daub-lined storage pit, a<br />

large clay-daubed storage basket, or even a claylined<br />

oven or hearth. It is possible that refitting<br />

these daub fragments may shed some light on the<br />

original structure they formed part <strong>of</strong>.<br />

4.3.2. Shell, Worked Bone and other utilized<br />

Fauna<br />

Several kinds <strong>of</strong> organic materials were<br />

utilized to make tools. As noted previously (Robb<br />

1998), a small fossil shark’s tooth from Trench 3<br />

may have been kept as a curiosity or used as a tool<br />

for some purpose such as incising unfired pottery.<br />

Similarly, two fragments <strong>of</strong> boar’s tusk from the<br />

upper strata <strong>of</strong> Trench 1 may have been used for<br />

tools or ornaments rather than simply representing<br />

culinary debris; one may be artifically smoothed at<br />

the end.<br />

One small marine shell, from a clam-like<br />

bivalve, was found in Stratum II <strong>of</strong> Trench 1. Its<br />

edges were artificially smoothed and rounded all<br />

around, and a small hole at its apex may have been<br />

naturally or artificially created. This example<br />

closely resembles a worked shell found in 1998 in<br />

the same stratum, as well as an unworked shell from<br />

the same stratum and a burnt shell fragment from<br />

Stratum IV. As these show, small marine shells<br />

were probably commonly used as tools for<br />

smoothing or scraping s<strong>of</strong>t materials.<br />

One worked bone implement was found<br />

(Figure 29). This was a long bone, probably a<br />

radius or metapodial from a sheep-sized animal,<br />

which had been split lengthwise and smoothed all<br />

around. The non-articular end was sharpened to a<br />

point, now broken, and the edges <strong>of</strong> the fragment<br />

were rounded. The result was a small awl <strong>of</strong> a type<br />

common in Neolithic sites. Its overall dimensions<br />

were 37 mm long by 10 mm wide by 8 mm thick.


4.3.3. Ground and Polished Stone<br />

Excavations to date have found between<br />

ten and twenty artifacts <strong>of</strong> ground and polished<br />

stone. The exact number is difficult to determine,<br />

since some pieces are clearly modified by humans,<br />

while others pieces consist <strong>of</strong> stone not found<br />

naturally on the site but do not display clear traces<br />

<strong>of</strong> working or shaping. It should also be noted that<br />

many <strong>of</strong> these are difficult for excavators to identify<br />

and a certain number <strong>of</strong> ground stone artifacts,<br />

particularly fragments or ones not obviously<br />

modified, may have been discarded accidentally<br />

during excavation.<br />

Non-flaked stone artifacts fall into several<br />

categories:<br />

• polished stone axes and other tools, and<br />

possible fragments from working or damaging<br />

them<br />

• fragments <strong>of</strong> large grinding stones, <strong>of</strong><br />

limestone or sandstone, with clearly worn<br />

surfaces<br />

• fragments <strong>of</strong> large metamorphic or igneous<br />

pebbles, sometimes showing worked facets or<br />

surfaces<br />

• small pebbles <strong>of</strong> unusual stone without any<br />

clear signs <strong>of</strong> human modification<br />

Polished stone axes and other tools. To<br />

date, two polished stone axes from Umbro have<br />

been found. The first was a broken fragment from<br />

the pointed (butt) end <strong>of</strong> a small, greenish-black<br />

axe or adze tool which was found in Stratum II <strong>of</strong><br />

Trench I in 1998 (see Robb 1998 for further detail).<br />

Excavations in 1999 uncovered a polished<br />

stone axe in Stratum II <strong>of</strong> Trench I (-10n/-<br />

36e)(Figure 30). Because <strong>of</strong> its position close to the<br />

surface (120-130 cm below datum, 10-20 cm<br />

below surface, in loose, sloping sediments), it is<br />

almost sure not to be in original depositional<br />

context. The axe is made <strong>of</strong> blackish amphibolite<br />

originally from the Sila (P. Barrier, pers. comm.). It<br />

measures 85 mm long by 47 mm high by 36 mm<br />

wide, and weighs 254 grams. The axe has an<br />

unusually round, stocky or barrel-shaped form<br />

which is unlike most published examples <strong>of</strong><br />

prehistoric axes from this region. Most <strong>of</strong> the body<br />

is highly polished and glossy. However, the small<br />

(butt) end is blunted with post-polishing damage <strong>of</strong><br />

some kind, and the large (blade) end is in the<br />

process <strong>of</strong> being reworked. The blade end has been<br />

broken <strong>of</strong>f flat to form a cylindrical end; a diagonal<br />

spall has been flaked <strong>of</strong>f this flat surface and<br />

partially smoothed, possibly as a beginning <strong>of</strong> refashioning<br />

a point to it; and in a band around the<br />

36<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

end on the sides <strong>of</strong> the axe the polished surface has<br />

been removed by pecking with a hard stone. The<br />

most likely interpretation seems that this piece was<br />

originally a much larger axe which was in the<br />

process <strong>of</strong> being remodelled into a smaller axe<br />

when it was discarded. The thickness, too great for<br />

a small axe, and the highly polished surface remain<br />

from the larger axe, which has been trimmed to a<br />

new, shorter length and was just starting to be<br />

thinned and re-pointed.<br />

Both <strong>of</strong> these axes are reasonably similar<br />

to other axes known from the area, for instance the<br />

three from loc. Cavalli near <strong>Bova</strong> Superiore in the<br />

Bruno Casile collection (examined in 1997 by kind<br />

permission <strong>of</strong> Sig. B. Casile). Unfortunately, such<br />

axes could have been used in any period from the<br />

Early Neolithic through the Bronze Age, and the<br />

relatively high stratigraphic positions <strong>of</strong> both<br />

Umbro examples does little to resolve their dating<br />

(see above). Polished stone axe manufacture may<br />

also be evident in the form <strong>of</strong> occasional thin spalls<br />

<strong>of</strong> hard stones alien to the site, whether worked or<br />

not.<br />

A more unusual polished stone artifact was<br />

the replica stone axe (Figure 30). Found close to<br />

the surface in Trench 1 (-10n/-36e, surface to 120<br />

cm), it cannot be considered in a secure<br />

stratigraphic context. It consists <strong>of</strong> a small, shaped<br />

and polished axe made <strong>of</strong> a locally available,<br />

streaky white, black and brown phyllite with mica<br />

(P. Barrier, pers. comm.). The surface is uniformly<br />

smooth except where small areas are broken away.<br />

The blade end is sharpened, but displays no use<br />

damage. The axe measures 4.5 mm thick, 10 mm<br />

high, and at least 17 mm long (part is missing). The<br />

stone is s<strong>of</strong>t and has strong cleavage planes, which<br />

would rule out the possibility <strong>of</strong> using it as a chisel<br />

or other small hand tool. Rather, it appears to be<br />

simply a replica axe, perhaps for use as a toy or<br />

amulet. Tiny replica axes or adzes are not unknown<br />

in the Southern Italian Neolithic, though most<br />

reported examples seem to be greenstone.<br />

Large grinding stones. Several stone<br />

fragments represent large querns or grinding stones<br />

made from relatively s<strong>of</strong>t sedimentary rock. One<br />

flat chunk <strong>of</strong> sandstone (135 x 112 x 48 mm, from<br />

Stratum II in the north end <strong>of</strong> Trench 1) had been<br />

shaped into an oval form; one flat face <strong>of</strong> it had<br />

been ground very smooth, while the other face and<br />

the side had been left rough. This fragment is<br />

probably a large piece <strong>of</strong> a two-handed rubber or<br />

mano. A second flat piece <strong>of</strong> sandstone with one<br />

very well-worn side (41 x 35 x 31, from the same<br />

general context) is probably part <strong>of</strong> a similar tool.<br />

Large querns are known from Umbro from<br />

only two fragments. This is a piece <strong>of</strong> a limestone<br />

slab 120 mm by 110 mm and 41 mm thick. One


edge <strong>of</strong> the fragment is raised with a lip ca. 1 cm<br />

high, so that the piece forms part <strong>of</strong> a very shallow<br />

basin. The top surface is smoothed. This piece<br />

probably formed part <strong>of</strong> a large flat quern or metate<br />

<strong>of</strong> a type common in Italian Neolithic villages. A<br />

second flat limestone fragment without a lip but <strong>of</strong><br />

similar dimensions (90 x 85 x 27) and smoothness<br />

probably comes from the same or a similar quern.<br />

Note that these two fragments were found<br />

in Trench 6, in an Early Bronze Age context. Two<br />

granite pebbles were also found in this context, one<br />

<strong>of</strong> which had a smooth facet and the other <strong>of</strong> which<br />

had possible ochre stains. It is possible that these<br />

are genuine EBA querns and grinding stones<br />

deposited in either a ritual or a domestic context;<br />

alternatively, they may be Neolithic tools re-used as<br />

building stone during the Bronze Age.<br />

Raw materials for these tools could have<br />

come from the site <strong>of</strong> Umbro itself, for limestone,<br />

or from nearby sandstone outcrops.<br />

Large pebbles. A third class <strong>of</strong> ground<br />

stone tools consisted <strong>of</strong> pebbles <strong>of</strong> granite and other<br />

non-local rock. These tools were identifiable in<br />

some cases by their raw material, and in other cases<br />

by evident modification. A total <strong>of</strong> eight were<br />

collected, two from Trench 6 and six from Trench<br />

1.<br />

• granite pebble with two grinding facets (86 x<br />

46 x 31 mm, broken; Bag 1066)<br />

• granite pebble with one side ground smooth<br />

(88 x 88 x 44 mm, broken; Bag 1121)<br />

• granite pebble with several smooth grinding<br />

facets and a slight pink/red stain, perhaps from<br />

ochre (60 x 41 x 45 mm, broken; Bag 1134)<br />

• pebble <strong>of</strong> unidentified metamorphic stone with<br />

several oblique grinding facets (76 x 58 x 37<br />

mm, broken; Bag 1135)<br />

• granite pebble with one side probably<br />

artificially smoothed (57 x 40 x 37 mm,<br />

broken; Bag 1149)<br />

• pebble <strong>of</strong> unidentified stone with one side<br />

probably artificially smoothed (69 x 58 x 39<br />

mm, broken; Bag 1149)<br />

• granite pebble with smoothed surfaces (79 x 58<br />

x 35 mm; Bag 1160, from Trench 6)<br />

• granite pebble without visibly smoothed facets<br />

but with possible ochre stain (68 x 56 x 45<br />

mm; Bag 1150, from Trench 6)<br />

It is likely that other examples were discarded<br />

inadvertently by excavators.<br />

The raw material for these pebble tools<br />

most probably came from the sandstone/<br />

37<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

conglomerate outcrops 200 meters north <strong>of</strong> Umbro.<br />

Their function was clearly grinding, as<br />

demonstrated by the flat facets worn on many <strong>of</strong><br />

their surfaces, and they were evidently curated and<br />

reused for some time, to judge from the multiple<br />

grinding facets many display. They may not have<br />

been the normal tools for grinding grain, and at<br />

least some may have been used to crush and grind<br />

ochre pigments, as several display reddish stains. It<br />

is not clear why most <strong>of</strong> them were broken; perhaps<br />

they were also used as hammerstones.<br />

Small pebbles. Finally, a number <strong>of</strong> small<br />

pebbles were found which were imported into the<br />

site and probably used as expedient tools. These<br />

include six examples from Trench 1 (Bag 1010;<br />

Bag 1136, 2 examples; Bag 1161; Bag 1211; Bag<br />

1216). Only one shows a possible smoothed facet<br />

(Bag 1211); the rest are identified solely by their<br />

non-local raw material. Other examples have<br />

probably been discarded by excavators. These<br />

pebbles are usually found unbroken and their<br />

maximum dimension is usually less than 50 mm.<br />

They may have been used for some purpose such as<br />

smoothing and burnishing unfired vessels.<br />

4.3.4. Ochre<br />

To date, eight small fragments <strong>of</strong> ochre<br />

have been found. All come from Trench 1, where<br />

they are found throughout the stratigraphy in all<br />

levels. The largest fragment is 21 mm in diameter,<br />

but most are smaller than 5 mm. They display no<br />

form or wear facets, but have clearly been imported<br />

to the site by humans for use as a pigment. This<br />

interpretation is corroborated by ochre stains on<br />

grinding stones (see above) and the use <strong>of</strong> red<br />

pigment to color impressions in vessels.<br />

It is worth noting that several levels in<br />

Trench 1 displayed slight concentrations <strong>of</strong> small,<br />

chalklike nodules. These were not collected, as they<br />

were s<strong>of</strong>t, disintegrating, and at times difficult to<br />

recognize. Such white lumps may represent a<br />

natural soil concretion, but they may also represent<br />

the remnants <strong>of</strong> materials used to make a white<br />

paste for encrusting impressed pottery.<br />

4.4. Faunal Remains and Human Remains<br />

Faunal Remains. The 1999 excavations<br />

recovered 892 fragments <strong>of</strong> animal bone; <strong>of</strong> these,<br />

5 came from Trench 6, 3 from Trench 7, and all the<br />

rest from Trench 1. Counting material from the<br />

1998 excavations, this brings the total faunal<br />

remains from Trench 1 to 1340 pieces (Table 2).<br />

This body <strong>of</strong> material will be studied by a trained<br />

faunal analyst following the 2000 field season. The<br />

following preliminary comments should be


understood as based on rapid impressions in the<br />

field by people untrained in animal bone<br />

identification.<br />

Taphonomically, the animal bone from<br />

Umbro is remarkable for its poor preservation.<br />

Most pieces are less than one centimeter in<br />

maximum dimension and are probably not<br />

identifiable. The most common identifiable pieces<br />

will almost certainly be tooth fragments. This<br />

marked degradation is not typical <strong>of</strong> extensive<br />

Neolithic village sites, in my experience, and serves<br />

to underscore the particular nature <strong>of</strong> Umbro. It<br />

probably derives from a very long-term use <strong>of</strong> a<br />

very restricted habitation area. It may also have<br />

something to do with the particular use <strong>of</strong> animal<br />

resources or food preparation or consumption at the<br />

site.<br />

Between one third and one quarter <strong>of</strong> all<br />

fragments are burnt, usually to a shade between<br />

black and white. Burning is found on bones from<br />

all strata <strong>of</strong> Trench 1 down to the earliest levels.<br />

This indicates exposure to extreme heat, as opposed<br />

to a brownish color which can indicate exposure to<br />

lesser temperatures. Intense burning on a par with<br />

cremation seems hardly compatible with thermal<br />

alteration in the process <strong>of</strong> cooking. Instead, it<br />

suggests that the sediments around Umbro were<br />

intensely heated periodically or repeatedly. Such<br />

intense burning no doubt contributed to the high<br />

fragmentation <strong>of</strong> the faunal remains.<br />

As an informal impression, the commonest<br />

kind <strong>of</strong> animal at the site seems to have been a<br />

small herbivore, most likely sheep and/or goat,<br />

though at this stage deer cannot be excluded.<br />

Several bones from very young individuals<br />

(lambs?) were found; these may prove informative<br />

on the season <strong>of</strong> occupation <strong>of</strong> the site. Pig has<br />

been identified as definitely present from several<br />

pieces submitted to faunal analysts (U. Albarella,<br />

pers. comm.; D. Serjeantson, pers. comm.), and<br />

teeth resembling pig are moderately common. It is<br />

unclear whether the pig would have been<br />

domesticated, wild boar, or feral. Cattle are<br />

probably represented by several fragments, but<br />

seem very uncommon. One fragment <strong>of</strong> cattle bone<br />

appears to be a nearly complete metapodial with<br />

numerous cut marks.<br />

Other animals present include bird <strong>of</strong><br />

some kind, snake and rodents. It is likely that some<br />

snake and rodent bones are intrusive modern<br />

specimens, especially as at least snakes now inhabit<br />

the site. Fish bones appear to be completely absent.<br />

Human Remains. In contrast to the 1998<br />

excavations which yielded a large assemblage <strong>of</strong><br />

disarticulated bone from Trench 3 (which later<br />

proved to be probably Roman in date), the 1999<br />

38<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

excavations yielded only three human skeletal<br />

fragments. All came from Stratum II in Trench 1.<br />

One consisted <strong>of</strong> a proximal pedal phalange<br />

fragment. The other two consisted <strong>of</strong> small cranial<br />

vault fragments whose precise location on the skull<br />

could not be determined. Like the four fragments <strong>of</strong><br />

human bones and teeth found in Trench 1 in 1998,<br />

these seem to be isolated fragments probably<br />

transported in from deposits higher up the southern<br />

slopes <strong>of</strong> the site and without any particular<br />

significance for the function <strong>of</strong> this part <strong>of</strong> the site.<br />

Supporting this interpretation is the fact that all<br />

seven fragments were found in Stratum II, a<br />

relatively high stratum probably representing mixed<br />

deposits.<br />

4.5. Paleobotanical Remains (<strong>Marina</strong> Ciaraldi)<br />

In 1999, we began a flotation program.<br />

Flotation samples were collected from all the<br />

features in Trench 6, for which the entire fill was<br />

bagged. This yielded four samples (one each from<br />

the three pots and one from the surrounding soil<br />

matrix). Samples were also taken from four<br />

representative levels in Trench 1, for which 20 liter<br />

samples were collected. For consistency, all Trench<br />

1 samples were taken from the same grid square (-<br />

7n/-36e); one sample was taken from each <strong>of</strong> strata<br />

II-V. Finally, one sample was taken from a feature<br />

in Stratum IV <strong>of</strong> Trench 1 (–8n/-36e). Flotation<br />

samples were first sifted in 5 mm mesh, like all<br />

excavated soil, and the finds were recorded and<br />

reunited with other finds from the bag. The samples<br />

were then floated by hand using basins and hand<br />

strainers. The coarse fraction was collected using 1<br />

mm nylon mesh, and the fine fraction was collected<br />

using a .5 mm nylon mesh. These fractions were<br />

then dried in fine mesh bags. The residue and the<br />

flotation (fine) fraction were both sorted and<br />

identified for seed and charcoal analysis by Dr.<br />

<strong>Marina</strong> Ciaraldi (Birmingham University Field<br />

<strong>Archaeological</strong> Unit). In the case <strong>of</strong> the flotation<br />

fraction, a standard low power stereomicroscope<br />

was used. Identifications were made using a<br />

reference collection <strong>of</strong> modern material and various<br />

identification atlases.<br />

None <strong>of</strong> the four samples from the Trench<br />

6 (Early Bronze Age) pots and feature fill<br />

contained organic remains. This is unsurprising as<br />

most <strong>of</strong> these samples were less than a liter in<br />

volume, and the largest was less than five liters.<br />

The remaining five samples all yielded<br />

botanical remains (Table 5). Sample 9 represents<br />

Stratum II, possibly mixed later Neolithic deposits;<br />

Sample 7 represents Stratum III (Diana); and<br />

Samples 5, 8, and 6 represent Stentinello deposits


from Stratum IV, Stratum V and a feature<br />

respectively.<br />

Many modern rootlets and some modern<br />

seeds were observed in sample 9 and, to a minor<br />

degree, in sample 7. The remaining three samples<br />

presented a very similar sandy matrix. All the<br />

samples contained some charcoal (although in very<br />

small quantity) and all samples but N.6 contains<br />

some small bones. The general preservation <strong>of</strong> the<br />

charred plant remains was generally poor. The<br />

cereal grains had clearly undergone carbonization<br />

under intense heat and they presented a damaged<br />

surface and numerous holes. The same applied to<br />

some <strong>of</strong> the chaff elements. The identification to<br />

species level was therefore difficult and in many<br />

cases impossible. Most <strong>of</strong> the cereal grains<br />

recovered were fragmented and distorted and they<br />

were assigned to the wide category <strong>of</strong> cereals.<br />

In some cases it was possible to attribute<br />

some <strong>of</strong> the grains to barley (Hordeum vulgare L.).<br />

However, even in these cases it was impossible to<br />

observe details such as the symmetry <strong>of</strong> the<br />

embryo, the presence <strong>of</strong> hull on the surface <strong>of</strong> the<br />

grain or eventual signs <strong>of</strong> germination. It was<br />

therefore impossible to establish whether these<br />

grains belonged to the six-row or to the two-row<br />

variety. In two cases some barley grains were<br />

identified simply as Hordeum sp. because they<br />

were rather small and they could have been either<br />

tail grains (underdeveloped grains) <strong>of</strong> cultivated<br />

barley or a species <strong>of</strong> barley.<br />

Only two wheat grains were found and<br />

only one was identified as bread wheat (Triticum<br />

aestivum s.l.). The diagnostic features used to<br />

identify this grain were its roundness and its curved<br />

bottom pr<strong>of</strong>ile. Most <strong>of</strong> the charred chaff found was<br />

damaged too and although the presence <strong>of</strong> emmer<br />

(Triticum dicoccum Schank.) was confirmed by the<br />

identification <strong>of</strong> some <strong>of</strong> the better preserved glume<br />

bases and forks, it was impossible to establish with<br />

certainty the presence <strong>of</strong> einkorn (Triticum<br />

monococcum L.).<br />

None <strong>of</strong> the weed seeds found is particular<br />

informative in the interpretation <strong>of</strong> the assemblage<br />

mainly because they could not be identified to<br />

species level. One possible olive pit was recovered<br />

from a good Early Neolithic context and identified<br />

in the field (identification by L. Foxhall); however,<br />

this did not form part <strong>of</strong> the material available for<br />

identification by a trained paleobotanist, and the<br />

identification must be considered tentative. If it is<br />

indeed an olive pit, it is likely to represent wild<br />

olive and may have been used for animal fodder (L.<br />

Foxhall, pers. comm.).<br />

The paucity <strong>of</strong> the plant remains recovered<br />

does not allow to make any general statment on the<br />

39<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

economy <strong>of</strong> the site. However, considering the<br />

general dearth <strong>of</strong> data from the Neolithic <strong>of</strong> the<br />

area, the plant material from Umbro provides an<br />

important contribution to the more general<br />

reconstruction <strong>of</strong> the Neolithic economy <strong>of</strong><br />

Calabria.<br />

Overall the plant assemblage from Umbro<br />

represents a typical assemblage from a Southern<br />

Italian Neolithic site, with the simultaneous<br />

presence <strong>of</strong> three different cereals: barley, emmer<br />

(and einkorn?) and a free-threshing wheat. This<br />

group <strong>of</strong> cereals represents the typical Neolithic<br />

“package” <strong>of</strong> Southern Europe (Zohary and Hopf<br />

1993). The absence <strong>of</strong> pulses from the plant<br />

assemblage may be the consequence <strong>of</strong> the<br />

generally poor preservation <strong>of</strong> the plant material.<br />

The agricultural component <strong>of</strong> the economy may<br />

also be attested by the common finds <strong>of</strong> grinding<br />

stone fragments (see above).<br />

There are only two Neolithic sites <strong>of</strong> the<br />

Stentinello facies from Calabria whose plant<br />

assemblage has been studied: Capo Alfiere,<br />

Crotone (CZ) (Costantini 1988) on the Ionic coast<br />

and Acconia area C, Curingia (CZ) (Ammerman<br />

1987) on the Tyrrenic coast. The information on the<br />

plant assemblage from Acconia is very limited and<br />

it only attests the presence <strong>of</strong> barley (Costantini and<br />

Stancanelli 1994). For the plant assemblage from<br />

Capo Alfiere we only have a preliminarily report<br />

but there is evidence <strong>of</strong> a higher number <strong>of</strong> species.<br />

The plant assemblage from Capo Alfiere reflects<br />

closely the findings from Umbro although there is<br />

no reference to the presence <strong>of</strong> chaff at Capo<br />

Alfiere (Costantini and Stancanelli 1994).<br />

The presence <strong>of</strong> grains, chaff and a few<br />

weeds (possibly associated with the cereal crops) is<br />

an important indication that part <strong>of</strong> the crop<br />

processing took place on site. The crop processing<br />

may have happened at different levels. It could<br />

have taken place as a large-scale crop cleaning<br />

directly on site or it may just have taken place as<br />

daily cleaning <strong>of</strong> cereals to be eaten. A higher<br />

density <strong>of</strong> charred plant remains was recovered<br />

from the feature fill (Sample 6, T1 8n.36e –230-<br />

240). Its composition was very similar to the rest <strong>of</strong><br />

the plant assemblage. This could perhaps be<br />

interpreted as evidence <strong>of</strong> the fact that crop<br />

cleaning took place on a daily basis and that the<br />

plant assemblage recovered from the feature fill<br />

reflects this activity.<br />

It is evident from the data that there is no<br />

clear differentiation between the assemblages from<br />

the different strata and occupational periods. If in<br />

some cases this can be due to a certain degree <strong>of</strong> remixing<br />

<strong>of</strong> the layers (especially for Stratum II and<br />

III), on the other hand this might be interpreted as a<br />

sign <strong>of</strong> continuation between the Stentinello and


Diana occupation levels. This seems in agreement<br />

with the hypothesis that these two kinds <strong>of</strong> ware<br />

might have been in use at the same time over a<br />

certain number <strong>of</strong> centuries (Whitehouse 1986).<br />

However the data are too scarce to be used with<br />

confidence in confirming this statement.<br />

40<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


San Pasquale is a Neolithic site located on<br />

a natural clay terrace at the eastern margin <strong>of</strong> the<br />

San Pasquale valley. It was found by S. Stranges<br />

and L. Saccà and first surveyed by the <strong>Bova</strong> <strong>Marina</strong><br />

<strong>Archaeological</strong> <strong>Project</strong> in 1997 (see survey site<br />

catalog, above, in which it is referred to as<br />

Canturatta A, Area 2). Both Stranges and Saccà and<br />

our survey have found pottery to include mostly<br />

small, highly eroded fragments with a few<br />

diagnostic impressed or Stentinello pieces; there<br />

are also some fragments which may be later<br />

prehistoric and some apparently Roman fragments<br />

which may be associated with Site 2.<br />

The natural setting <strong>of</strong> the site would have<br />

been quite attractive for Neolithic settlement. It is<br />

located about 100 m from the present day coastline<br />

and about 30 m above sea level, overlooking the<br />

broad, level valley <strong>of</strong> the Torrente di S. Pasquale.<br />

The terrace itself consists <strong>of</strong> pure clayey and sandy<br />

deposits, as its eroding margins show. A deep<br />

ravine divides it into an eastern and a western part.<br />

Like most such terraces, it has now been built up<br />

with a cascade <strong>of</strong> small agricultural terraces, but it<br />

seems likely that this reflects the underlying shape<br />

<strong>of</strong> the hill substantially. In prehistoric times, the<br />

torrente may have had marshy areas at its mouth,<br />

and the valley would have probably been wooded.<br />

Plenty <strong>of</strong> level ground would have been available<br />

for building and farming, as well as marine<br />

resources and a variety <strong>of</strong> montane resources from<br />

the Agrillei ridge just behind the site. The ridge<br />

would have also sheltered the site from northern<br />

and eastern winds. In addition, there may well have<br />

been springs available at points around the base <strong>of</strong><br />

the Agrillei ridge; a large clump <strong>of</strong> reeds near the<br />

eastern end <strong>of</strong> the terrace suggests some ground<br />

moisture, and a spring is still extant at a similar<br />

stratigraphic level below the ridge on the western<br />

side <strong>of</strong> the valley.<br />

Such a location – on a low terrace<br />

commanding ecotonal resources – is <strong>of</strong>ten<br />

considered virtually stereotypical <strong>of</strong> Neolithic<br />

villages, based on areas such as the Materano and<br />

the Tavoliere (Jarman and Webley 1975; Jones<br />

1987; Tinè 1983). Hence, if we suppose that the<br />

Calabrian Neolithic was based around large villages<br />

as in other areas, San Pasquale seems an obvious<br />

area to look for one. Thus, we were especially<br />

interested in test-excavating San Pasquale to<br />

investigate sites complementary to Umbro, which is<br />

clearly not a large, open village.<br />

At present, the area <strong>of</strong> the site is used for<br />

several kinds <strong>of</strong> agricultural production. On the<br />

5. TEST EXCAVATIONS AT SAN PASQUALE<br />

41<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

western part, the terrace is in derelict scrub, and<br />

had recently been burnt over at the time <strong>of</strong><br />

excavation; hence surface visibility was excellent.<br />

Across the ravine on the broader eastern part, there<br />

was an active and fenced vineyard and, further east,<br />

a grove <strong>of</strong> olive and almond trees with some<br />

scrubby vegetation between them.<br />

As a preliminary to test excavations, we<br />

examined current ground surfaces as well as<br />

interviewing past surveyors (especially D. Yoon<br />

and L. Saccà) as to the location <strong>of</strong> artifact<br />

concentrations. It is clear that some cultural<br />

materials are found throughout the terrace, but that<br />

there are at least two distinct concentrations. One is<br />

located in the center <strong>of</strong> the western part <strong>of</strong> the<br />

terrace. Here we found, for instance, scattered<br />

prehistoric and historic sherds, all small eroded<br />

fragments. A second is located along the southern<br />

margin <strong>of</strong> the center <strong>of</strong> the eastern plateau, between<br />

a large outcropping boulder (collected as “Location<br />

B”) and the nearby olive trees (collected as<br />

“Location A”). Finds here seemed to be<br />

predominantly Roman or modern, though at least<br />

one flake <strong>of</strong> obsidian suggests prehistoric<br />

occupation.<br />

We decided to excavate test pits on the<br />

western part <strong>of</strong> the terrace in the center <strong>of</strong> the<br />

concentration <strong>of</strong> sherds there (Figure 31). Two one<br />

by one meter test pits were excavated, Trench 1 and<br />

Trench 2 (Figure 32, Figure 33). The two pits were<br />

aligned to a hypothetical north-south grid and<br />

spaced 16 meters apart. They were excavated in ten<br />

centimeter arbitrary levels. Work was carried out<br />

with a crew <strong>of</strong> 8 on September 9 th .<br />

The results were disappointing, but at least<br />

they were clear. In both pits, the ground surface<br />

was littered with small, tabular gravel, consisting<br />

largely <strong>of</strong> schist fragments evidently originating<br />

from the slopes <strong>of</strong> Agrillei just 40-60 meters to the<br />

northeast. When this was removed, the soil rapidly<br />

turned to a brownish sandy clay which contained<br />

virtually no archaeological materials. As the<br />

trenches grew deeper, the soil grew yellower and<br />

both stones and bulbs and roots from surface<br />

vegetation became fewer and fewer. Both trenches<br />

were stopped between 40 and 50 cm deep, when it<br />

was clear that the soil below the surface was<br />

essentially sterile and was rapidly turning identical<br />

to the 20+ meter thick deposit <strong>of</strong> clay visible in the<br />

eroding banks <strong>of</strong> the terrace. Both trenches were<br />

backfilled.


While these excavations were taking<br />

place, three crew cleaned pr<strong>of</strong>iles in eroding places<br />

at four locations around the edges <strong>of</strong> the terrace. In<br />

all four places, with pr<strong>of</strong>iles up to three meters<br />

deep, we encountered stratigraphy identical to that<br />

in the test trenches: a litter <strong>of</strong> stone fragments on<br />

the surface, followed by a thin transition to sterile<br />

geological deposits.<br />

The main, and apparently indisputable,<br />

conclusion is that the archaeological deposits in this<br />

particular place are entirely superficial. A<br />

suggestion was made by L. Saccà that farmers<br />

earlier in the century, to improve the quality <strong>of</strong> land<br />

for vines, dumped gravel into the field; few remains<br />

were visible on the surface because the real site lay<br />

buried a meter or more deep. However, this concept<br />

does not account for a number <strong>of</strong> facts: some<br />

artifacts are indeed found on the surface, the scatter<br />

<strong>of</strong> stones is entirely superficial rather than a thick<br />

stratum well mixed with the soil; and nowhere in<br />

exposed sections is there any evidence <strong>of</strong> a buried<br />

site outcropping at any depth.<br />

Instead, the most likely interpretation is<br />

that both the sheet <strong>of</strong> thin schist fragments and the<br />

scatter <strong>of</strong> artifacts identified as the site have washed<br />

in from the ridge to the northeast. This is the<br />

geological source <strong>of</strong> the schist, which has nothing<br />

to do with the s<strong>of</strong>t sedimentary deposits <strong>of</strong> the<br />

terrace. It would account for the apparent mixing <strong>of</strong><br />

the fragments, and would also explain their highly<br />

fragmented and eroded state, which is so severe that<br />

virtually none <strong>of</strong> them retains its original surface.<br />

The concentration <strong>of</strong> both stones and sherds on the<br />

surface <strong>of</strong> the field where we excavated can be<br />

explained through repeated cycles <strong>of</strong> alluviation<br />

and deflation.<br />

We thus concluded that there is a genuine<br />

Neolithic site somewhere in the vicinity <strong>of</strong> the area<br />

excavated, probably at no great distance. Although<br />

some <strong>of</strong> the relevant areas have been fieldwalked<br />

and do indeed yield low-level concentrations <strong>of</strong><br />

both prehistoric and historic material, it is unclear<br />

whether these represent the real site or are also<br />

slopewash. The task for the future is to walk and rewalk<br />

all areas in the environs, including some steep<br />

and thickly brushy areas, to check where the site<br />

may have been.<br />

42<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


6. CONCLUSIONS AND FUTURE RESEARCH DIRECTIONS<br />

6.1. Umbro: Dating, Stratigraphy, and Site<br />

Function<br />

The dating and stratigraphy <strong>of</strong> Umbro are<br />

gradually becoming clearer, thanks to numerous<br />

small test excavations and radiocarbon dates. The<br />

site was first occupied in the early 6 th millennium<br />

BC by people using Stentinello wares. The area<br />

occupied then consisted <strong>of</strong> a small zone in the<br />

shelter <strong>of</strong> low, east-facing cliffs. Stentinello<br />

occupation appears to have continued, continuously<br />

or intermittently, at least through the middle <strong>of</strong> the<br />

5 th millennium BC and likely for longer, as Late<br />

Neolithic Diana wares are common. There also<br />

appears to have been some Copper Age occupation,<br />

although this is poorly defined. Following a lapse<br />

<strong>of</strong> perhaps a thousand years, the site was reutilized<br />

in the Early Bronze Age by people using ceramics<br />

akin to the Eastern Sicilian Rodi-Tindari-<br />

Vallelunga wares. Nothing is known <strong>of</strong> later Bronze<br />

Age or Iron Age use <strong>of</strong> the site, but a small Greek<br />

farmstead was occupied from Archaic times<br />

onward.<br />

The nature <strong>of</strong> occupation in each period is<br />

less clear and more interesting. In the Neolithic,<br />

Umbro was a very small site, which could have<br />

been occupied by perhaps five to ten people at<br />

most. Influenced by areas such as Puglia, the<br />

Materano, the Adriatic coastal strip and eastern<br />

Sicily, archaeologists have usually understood<br />

Neolithic settlement as centered around villages<br />

housing many families. If this settlement pattern<br />

was found in Calabria as well, then Umbro must be<br />

understood as a special function site <strong>of</strong> some kind.<br />

It would have been occupied by a fraction <strong>of</strong><br />

society performing some combination <strong>of</strong><br />

specialized tasks such as hunting, herding, making<br />

pottery, or conducting rituals. Such a group may<br />

have been a nuclear family, or it may have been<br />

recruited along age, gender, ritual or other lines.<br />

However, we must also consider the possibility that<br />

Neolithic settlement in Calabria did not follow the<br />

pattern established elsewhere; it could have been<br />

decentralized, without fixed, aggregated villages<br />

(cf. Whittle 1996). Evidence from Trench 1 gives<br />

some idea <strong>of</strong> the nature <strong>of</strong> activities carried on at<br />

the site. Pots were made, probably in the dry<br />

season; cereals were prepared for consumption;<br />

domestic animals were eaten; exotic materials were<br />

fashioned into axes and cutting tools.<br />

The function <strong>of</strong> Umbro in the Early<br />

Bronze Age is currently equally ambiguous; as<br />

known from limited preliminary excavations, it may<br />

43<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

have been a burial site, a ritual site or a domestic<br />

site.<br />

The function <strong>of</strong> Umbro is inextricably<br />

linked to use <strong>of</strong> the surrounding landscape. For<br />

both the Neolithic and the Early Bronze Age,<br />

understanding the function <strong>of</strong> the site depends on<br />

understanding other sites in the neighborhood –<br />

whether Neolithic villages existed nearby, and<br />

whether there were domestic or funerary Bronze<br />

Age sites nearby.<br />

6.2. The Landscape <strong>of</strong> <strong>Bova</strong> <strong>Marina</strong> from<br />

Neolithic through Recent Times<br />

General trends in settlement as known<br />

from the field survey have been discussed above<br />

and are only briefly summarized here. Prehistoric<br />

settlement is theoretically poorly understood for the<br />

Neolithic and not known for later periods.<br />

However, survey data shows early settlement by<br />

Greeks, or acculturation <strong>of</strong> native populations to<br />

Greek customs. Surprisingly, hellenization occurred<br />

early even at small, inland farmsteads as well as in<br />

cities such as Rhegion. Probably the densest<br />

settlement in antiquity took place in the later<br />

Roman period, between the second and fifth<br />

centuries A.D., when intensive occupation <strong>of</strong><br />

coastal plains and river valleys probably reflects<br />

villa-based intensive agriculture. Little is known <strong>of</strong><br />

the following early medieval period until the<br />

foundation <strong>of</strong> inland hilltop towns such as <strong>Bova</strong><br />

Superiore. Population rose again in historic times,<br />

with dense peasant land use bespoken by numerous<br />

now-abandoned farmhouses. Perhaps the final turn<br />

in recent settlement history is a general shift in the<br />

last century back to the coastlands with the gradual<br />

abandonment <strong>of</strong> traditional mountain villages.<br />

6.3 Future Work<br />

Future work includes several logical<br />

developments <strong>of</strong> this season’s researches.<br />

Our field survey plans include three<br />

primary objectives:<br />

(1) to continue survey in the comune <strong>of</strong><br />

<strong>Bova</strong> <strong>Marina</strong>, exploring especially inland<br />

territories.<br />

(2) to expand survey inland to the comune<br />

<strong>of</strong> <strong>Bova</strong> Superiore, in order to understand<br />

settlement in the entire range <strong>of</strong> environmental<br />

zones locally available and used in historical times.


Areas to target specifically include the hills<br />

between Umbro and <strong>Bova</strong> Superiore, the immediate<br />

surroundings <strong>of</strong> <strong>Bova</strong> Superiore, and the Campi di<br />

<strong>Bova</strong>, a small highland plain traditionally used as<br />

pastures for people based at <strong>Bova</strong>.<br />

(3) to begin a program <strong>of</strong> small test<br />

excavations in undated prehistoric sites, to check<br />

the presence <strong>of</strong> periods such as the Copper and<br />

Bronze Ages which are very difficult to identify<br />

from survey materials and to verify the presence <strong>of</strong><br />

Neolithic sites in the surroundings <strong>of</strong> Umbro. The<br />

targets for 2000 include a probable Bronze Age site<br />

at Limaca, about 200 meters north <strong>of</strong> Umbro, and<br />

two undated ceramic scatters about 300 meters<br />

southwest <strong>of</strong> Umbro.<br />

The excavation goals are straightforward.<br />

We hope to continue excavations in the Neolithic<br />

44<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

area, though on a limited scale. The specific goals<br />

will be to further define the chronology <strong>of</strong> the site,<br />

to ascertain the beginning <strong>of</strong> occupation at the base<br />

<strong>of</strong> the sequence near the cliff, and to recover further<br />

scientific samples (for example, for flotation). We<br />

hope to conduct substantial excavations in the Early<br />

Bronze Age area, primarily to understand the nature<br />

<strong>of</strong> the site and, if it turns out a burial site, to learn<br />

about EBA funerary ritual. Finally, we hope to<br />

conduct test excavations in the Greek site about<br />

200 meters south <strong>of</strong> Umbro. As noted earlier,<br />

excavation <strong>of</strong> a small, early inland Greek farmstead<br />

has the potential to yield much new information<br />

about social processes <strong>of</strong> acculturation in the early<br />

colonial period.


Ammerman, A.<br />

Bernabò Brea, L.<br />

1987 Ricenti contributi sul neolitico della<br />

Calabria. Atti, Riunione Scientifica<br />

dell'I.I.P.P. 26:333-349.<br />

1957 Sicily before the Greeks. Thames and<br />

Hudson, London.<br />

Bernabò Brea, L., and M. Cavalier<br />

Castellana, G.<br />

Ceci, M.<br />

1956 Civiltà preistoriche delle Isole Eolie e<br />

del territorio del Milazzo. Bullettino di<br />

Paletnologia Italiana 66:7-98.<br />

1960 Meligunìs Lipára. Volume I: La<br />

stazione preistorica della contrada Diana e<br />

la necropoli preistorica di Lipari.<br />

Pubblicazioni del Museo Eoliano di Lipari.<br />

S. F. Flaccovio, Palermo.<br />

1980 Meligunìs Lipára, Volume IV:<br />

l'acropoli di Lipari nella preistoria.<br />

Publications <strong>of</strong> the Museo Eolio. Flaccovio,<br />

Palermo.<br />

1996 La stipe votiva del Ciavolaro nel<br />

quadro del Bronzo Antico Siciliano.<br />

Assessorato regionale beni culturali<br />

ambieltali e pubblica istruzione, Agrigento.<br />

1992 Note sulla circolazione delle lucerne<br />

a Roma nell'VIII secolo: i contesti della<br />

Crypta Balbi. Archeologia Medievale<br />

19:749-766.<br />

Cocchi Genick, D.<br />

Costamagna, L.<br />

Costantini, L.<br />

1996 Manuale di Preistoria. III. Età del<br />

Rame. Octavo, Firenze.<br />

1991 La sinagoga di <strong>Bova</strong> <strong>Marina</strong> nel<br />

quadro degli insediamenti tardoantichi della<br />

costa jonica meridionale della Calabria.<br />

Mèlanges de l'école française di Rome 2.<br />

1988 Cereali e legumi del sito neolitico di<br />

Capo Alfiere, Unpublished manuscript.<br />

Costantini, L., and M. Stancanelli<br />

1994 La preistoria agricola dell' Italia<br />

centro-meridionale: il contributo delle<br />

indagini archeobotaniche. Origini 18:149-<br />

243.<br />

BIBLIOGRAPHY<br />

45<br />

Foley, R.<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

1981 A model <strong>of</strong> regional archaeological<br />

structure. Proceedings <strong>of</strong> the Prehistoric<br />

Society 47:1-17.<br />

Garcea, F., and D. Williams<br />

Gregg, S.<br />

1987 Appunti sulla produzione e<br />

circolazione delle lucerne nel napoletano tra<br />

VII e VIII secolo. Archeologia Medievale<br />

14:537-545.<br />

1988 Foragers and farmers: population<br />

interaction and agricultural expansion in<br />

prehistoric Europe. University <strong>of</strong> Chicago<br />

Press, Chicago.<br />

Higgs, E., and C. Vita-Finzi<br />

1972 Prehistoric economies: a territorial<br />

approach. In Papers in economic<br />

prehistory, edited by E. Higgs, pp. 27-36.<br />

Cambridge University Press, Cambridge.<br />

Jarman, M., and D. Webley<br />

1975 Settlement and land use in<br />

Capitanata, Italy. In Palaeoeconomy, edited<br />

by E. Higgs, pp. 177-231. Cambridge<br />

University Press, Cambridge.<br />

Jarman, M. G., G. Bailey, and H. Jarman<br />

Jones, G.B.D.<br />

Leighton, R.<br />

Maniscalco, L.<br />

1982 Early European agriculture.<br />

Cambridge University Press, Cambridge.<br />

1987 Apulia. Society <strong>of</strong> Antiquaries,<br />

London.<br />

1999 Sicily before history. Duckworth,<br />

London.<br />

1996 Early Bronze Age funerary ritual and<br />

architecture: monumnetal tombs at Santa<br />

Febronia. In Early Societies in Sicily, edited<br />

by R. Leighton, pp. 81-87. Accordia<br />

Research Centre, London.<br />

1999 The Sicilian Bronze Age pottery<br />

service. In Social Dynamics <strong>of</strong> the<br />

Prehistoric Central Mediterranean, edited<br />

by R. H. Tykot, J. Morter, and J. E. Robb,<br />

pp. 185-194. Accordia Research Centre,<br />

London.


Morter, J.<br />

O'Hare, G.<br />

Orsi, P.<br />

Robb, J.<br />

Robb, J. E.<br />

1992 Capo Alfiere and the Middle<br />

Neolithic period in eastern Calabria,<br />

Southern Italy. Unpublished Ph.D.<br />

Dissertation, University <strong>of</strong> Texas, Austin.<br />

1990 A preliminary study <strong>of</strong> polished stone<br />

artefacts in prehistoric southern Italy.<br />

Proceedings <strong>of</strong> the Prehistoric Society<br />

56:123-152.<br />

1890 Stazione neolitica di Stentinello.<br />

Bullettino di Paletnologia Italiana 16:177-<br />

200.<br />

1997 <strong>Bova</strong> <strong>Marina</strong> Field Survey:<br />

Preliminary Report, 1997 Season.<br />

<strong>Department</strong> <strong>of</strong> <strong>Archaeology</strong>, University <strong>of</strong><br />

Southampton, Southampton.<br />

1998 <strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong><br />

<strong>Project</strong>: Survey and Excavations at Umbro.<br />

<strong>Department</strong> <strong>of</strong> <strong>Archaeology</strong>, University <strong>of</strong><br />

Southampton, Southampton.<br />

Servizio Geologico dell'Italia<br />

Shaffer, G.<br />

Skeates, R.<br />

1970 Carta Geologica dell'Italia. Cassa<br />

per il Mezzogiorno. 1:25000.<br />

1976 Carta Geologica dell'Italia. IGM.<br />

1:500000.<br />

1985 Architectural resources and their<br />

effect on certain neolithic settlements in<br />

Southern Italy. In Papers in Italian<br />

<strong>Archaeology</strong> IV: the Cambridge<br />

conference, edited by C. Malone, and S.<br />

Stoddart, pp. 101-117. BAR International<br />

Series, 245. British <strong>Archaeological</strong> Reports,<br />

Oxford.<br />

1992 Thin-section analysis <strong>of</strong> Italian<br />

neolithic pottery. In Papers <strong>of</strong> the Fourth<br />

Conference <strong>of</strong> Italian <strong>Archaeology</strong>. Volume<br />

3: New developments in Italian<br />

archaeology, edited by E. Herring, R.<br />

46<br />

Stranges, S.<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

Whitehouse, and J. Wilkins, pp. 29-34.<br />

Accordia Research Center, London.<br />

1994 A radiocarbon date-list for prehistoric<br />

Italy (c. 46,400 BP - 2450 BP/400 cal. BC).<br />

In Radiocarbon dating and Italian<br />

prehistory, edited by R. Skeates, and R.<br />

Whitehouse, pp. 147-288. Accordia<br />

Specialist Studies on Italy, 3. Accordia<br />

Research Center, London.<br />

1992 Importante ritrovamento<br />

stentinelliano a <strong>Bova</strong> <strong>Marina</strong>. Calabria<br />

Sconosciuta 15:51-52.<br />

Stranges, S., and L. Saccà<br />

Tinè, S.<br />

1994 Nuove acquisizioni sulla preistoria<br />

nella Jonica Reggina. Calabria Sconosciuta<br />

17:19-22.<br />

1983 Passo di Corvo e la civiltà neolitica<br />

del Tavoliere. Sagep, Genova.<br />

1992 <strong>Bova</strong> Survey 1992. Istituto Italiano di<br />

Archeologia Sperimentale, Genova.<br />

Tozzi, C., and G. Tasca<br />

Tusa, S.<br />

Whitehouse, R.<br />

Whittle, A.<br />

1989 Ripa Tetta. Atti Convegno Nazionale<br />

sulla Preistoria, Protostoria e Storia della<br />

Daunia 11:39-54.<br />

1993 La Sicilia nella preistoria. 2nd ed.<br />

Sellerio, Palermo.<br />

1986 Siticulosa Apulia revisited. Antiquity<br />

60:36-44.<br />

1996 Neolithic Europe: the creation <strong>of</strong> new<br />

worlds. Cambridge University Press,<br />

Cambridge.<br />

Zohary, D., and M. Hopf<br />

1993 Domestication <strong>of</strong> plants in the Old<br />

World. Clarendon, Oxford.


Figures<br />

1. Map <strong>of</strong> areas surveyed and sites found, 1997-1999.<br />

2. Map <strong>of</strong> areas surveyed in the Umbro area.<br />

3. Map <strong>of</strong> areas surveyed in the Mazza area.<br />

4. Map <strong>of</strong> grid collection units at Mazza.<br />

5. GIS reconstruction <strong>of</strong> prehistoric land use.<br />

6. Umbro: site map.<br />

FIGURES AND TABLES<br />

47<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

7. Umbro: general view <strong>of</strong> cliffs with Neolithic site below, Bronze Age site above (from north).<br />

8. Trench 1 plan.<br />

9. Trench 1 section.<br />

10. Trench 1 general view at end <strong>of</strong> excavations.<br />

11. Trench 6 pr<strong>of</strong>ile <strong>of</strong> road cut before excavation <strong>of</strong> trench.<br />

12. Trench 6 plan <strong>of</strong> excavations.<br />

13. Trench 6 pr<strong>of</strong>ile <strong>of</strong> excavations.<br />

14. Trench 6 plan <strong>of</strong> pot group.<br />

15. Trench 6 general view <strong>of</strong> Bronze Age surface and rock fall.<br />

16. Trench 6 Bronze Age pots in situ.<br />

17. Trench 7 plan.<br />

18. Trench 7 pr<strong>of</strong>ile.<br />

19. Trench 7 general view.<br />

20. Clearing bedrock near peak <strong>of</strong> site.<br />

21. Area 2 general view (1998 excavations).<br />

22. Trench 3 general view (1998 excavations).<br />

23. Trench 4 general view (1998 excavations).<br />

24. Trench 5 general view (1998 excavations).<br />

25. Umbro: Stentinello pottery.<br />

26. Umbro: Diana pottery.<br />

27. Umbro: Other prehistoric pottery.<br />

28. Umbro: Early Bronze Age vessels from Trench 6 deposition.<br />

29. Umbro: miscellaneous tools: (a) stampini, (b) worked bone, (c) worked shell.<br />

30. Umbro: ground stone (a) axe, (b) miniature replica axe.<br />

31. San Pasquale general view <strong>of</strong> location <strong>of</strong> test pits.<br />

32. San Pasquale Trench 1 test pit.<br />

33. San Pasquale Trench 2 test pit.


Tables<br />

Table 1. Trench 1 finds and taphonomic characteristics by depth.<br />

Table 2. Summary <strong>of</strong> faunal assemblage from Trench 1.<br />

Table 3. Summary <strong>of</strong> lithic assemblage from Trench 1.<br />

Table 4. Summary <strong>of</strong> daub frequencies in Trench 1.<br />

Table 5. Paleobotanical samples.<br />

Table 6. Radiocarbon dates.<br />

48<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


1. Map <strong>of</strong> areas surveyed and sites found, 1997-1999.<br />

49<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


2. Map <strong>of</strong> areas surveyed in the Umbro area.<br />

50<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


3. Map <strong>of</strong> areas surveyed in the Mazza area.<br />

51<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


4. Map <strong>of</strong> grid collection units at Mazza.<br />

52<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


5. GIS reconstruction <strong>of</strong> prehistoric land use.<br />

53<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


6. Umbro: site map.<br />

54<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


55<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

7. Umbro: general view <strong>of</strong> cliffs with Neolithic site below, Bronze Age site above (from north).


8. Trench 1 plan.<br />

56<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


9. Trench 1 section.<br />

57<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


10. Trench 1 general view at end <strong>of</strong> excavations.<br />

58<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


11. Trench 6 pr<strong>of</strong>ile <strong>of</strong> road cut before excavation <strong>of</strong> trench.<br />

59<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


12. Trench 6 plan <strong>of</strong> excavations.<br />

60<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


13. Trench 6 pr<strong>of</strong>ile <strong>of</strong> excavations.<br />

61<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


14. Trench 6 plan <strong>of</strong> pot group.<br />

62<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


15. Trench 6 general view <strong>of</strong> Bronze Age surface and rock fall.<br />

63<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


16. Trench 6 Bronze Age pots in situ.<br />

64<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


17. Trench 7 plan.<br />

65<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


18. Trench 7 pr<strong>of</strong>ile.<br />

66<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


19. Trench 7 general view.<br />

67<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


20. Clearing bedrock near peak <strong>of</strong> site.<br />

68<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


21. Area 2 general view (1998 excavations).<br />

69<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


22. Trench 3 general view (1998 excavations).<br />

70<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


23. Trench 4 general view (1998 excavations).<br />

71<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


24. Trench 5 general view (1998 excavations).<br />

72<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


25. Umbro: Stentinello pottery.<br />

73<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


26. Umbro: Diana pottery.<br />

74<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


27. Umbro: Other prehistoric pottery.<br />

75<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


28. Umbro: Early Bronze Age vessels from Trench 6 deposition.<br />

76<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


29. Umbro: miscellaneous tools: (a) stampini, (b) worked bone, (c) worked sherd.<br />

77<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


30. Umbro: ground stone (a) axe, (b) miniature replica axe.<br />

78<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


31. San Pasquale general view <strong>of</strong> location <strong>of</strong> test pits.<br />

79<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


32. San Pasquale Trench 1 test pit.<br />

80<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


33. San Pasquale Trench 2 test pit.<br />

81<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


Table 1. Trench 1 finds and taphonomic characteristics by depth.<br />

Stratum Depth Ceramics Obsidian Flint Fauna Daub Ground<br />

stone<br />

82<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

% Fauna<br />

burnt<br />

sherd<br />

diameter<br />

(mean)<br />

obsidian<br />

length<br />

(mean)<br />

I 0-130 528 59 2 186 54 1 37 29.7 15.7<br />

II 130-160 1172 110 6 329 269 10 29 33.9 19.9<br />

III 160-190 710 106 7 338 361 8 19 32.6 17.2<br />

IV 190-230 388 71 2 343 330 0 17 32.6 19.0<br />

V* 230-250 99 28 0 135 80 1 20 26.1 17.7


Table 2. Summary <strong>of</strong> faunal assemblage from Trench 1.<br />

stratum depth unburned burned<br />

fragments fragments<br />

1 0-100 44 13<br />

1 0-120 8 6<br />

1 0-130 2 1<br />

1 0-147 3 1<br />

1 100-110 22 40<br />

1 110-120 16 7<br />

1 120-130 18<br />

1 135-150 4 1<br />

2 110-120 21 22<br />

2 120-130 27 9<br />

2 130-140 47 14<br />

2 140-150 50 26<br />

2 150-160 77 18<br />

2 160-170 13 5<br />

3 160-170 62 28<br />

3 170-180 82 18<br />

3 180-190 74 13<br />

3 190-200 49 3<br />

3 200-210 7 2<br />

4 200-210 69 17<br />

4 210-220 76 11<br />

4 220-230 83 12<br />

4 230-240 25 13<br />

4 240-250 30 7<br />

5 240-250 67 3<br />

5 250-260 24 19<br />

5 260-270 16 2<br />

5 270-280 1 3<br />

83<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


Table 3. Summary <strong>of</strong> lithic assemblage from Trench 1.<br />

84<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

stratum raw blade blade chunk core flake tool<br />

mat.<br />

fragment<br />

1 f 1 1<br />

1 o 4 5 5 1 42 2<br />

2 f 1 4 1<br />

2 o 3 12 12 4 72 7<br />

3 f 1 1 4 1<br />

3 o 3 20 12 1 63 7<br />

4 f 1 1<br />

4 o 3 10 7 1 48 2<br />

5 o 5 2 20 1


Table 4. Summary <strong>of</strong> daub frequencies from Trench 1.<br />

stratum depth Total<br />

fragments/level<br />

Average<br />

fragments/<br />

square<br />

1 0-100 8 2.66<br />

1 0-120 5 5<br />

1 0-147 3 3<br />

1 100-110 13 6.5<br />

1 110-120 9 9<br />

1 120-130 7 7<br />

1 135-150 9 9<br />

2 110-120 9 2.25<br />

2 120-130 9 2.25<br />

2 130-140 36 7.2<br />

2 140-150 58 5.8<br />

2 150-160 128 11.64<br />

2 160-170 29 29<br />

3 160-170 107 11.89<br />

3 170-180 91 10.11<br />

3 180-190 46 5.11<br />

3 190-200 76 12.67<br />

3 200-210 41 41<br />

4 200-210 129 16.13<br />

4 210-220 116 16.57<br />

4 220-230 40 10<br />

4 230-240 39 9.75<br />

4 240-250 6 3<br />

5 240-250 46 15.33<br />

5 250-260 34 17<br />

85<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999


Table 5. Paleobotanical samples.<br />

common<br />

name<br />

86<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

Sample N. 9 7 5 8 6<br />

Feature/Context T1 7n-36e<br />

(160-170)<br />

str.II<br />

T1 7n.36e<br />

str.III<br />

T1 7n.36e<br />

(230-240)<br />

Str IV<br />

T1 7n.36e<br />

(250-260)<br />

str.V<br />

T1 8n.37e<br />

(230-240)<br />

Sample vol.<br />

Seed/Litre<br />

20 20 20 20 2.5<br />

flot/res f f f f f<br />

Type <strong>of</strong> context stratum stratum stratum stratum feature fill<br />

Period/phase Neol Neol Neol Neol Neol<br />

Taxon body<br />

part<br />

pres.<br />

cereals Cerealia gr carb 4 6 10 6 6<br />

cereals Cerealia c.n. carb 1<br />

barley Hordeum cf.<br />

vulgare<br />

gr carb 6 1 2 4 3<br />

(wild?)<br />

barley<br />

Hordeum sp. gr carb 1 1<br />

heinkorn/ Triticum gl.b. carb 3 2 3 1 1<br />

emmer monococcum/<br />

dicoccum<br />

heinkorn/ Triticum f carb 1 1 1<br />

emmer monococcum/<br />

dicoccum<br />

emmer Triticum<br />

dicoccum<br />

gl.b. carb 1 1 1 2<br />

emmer Triticum<br />

dicoccum<br />

f carb 1<br />

bread Triticum aestivum gr carb 1<br />

wheat s.l.<br />

wheat Triticum sp. gr carb 1<br />

wheat Triticum sp. r.in. carb 1 3 2<br />

Atriplex sp. s carb 1<br />

vetch Vicia/Lathyrus s carb 1<br />

Trifolium/Medica<br />

go/ Melilotus<br />

s carb 1<br />

grape Vitis vinifera s mo 1<br />

grasses Gramineae (Poa<br />

type)<br />

s carb 1 1<br />

grasses Gramineae s carb 1<br />

modern<br />

seeds<br />

Charcoal carb 5fr 3fr 10fr 15fr 15fr<br />

eggs/coprolites carb x 3<br />

cyst 2 1 2<br />

small bones x x x x


Table 6. Radiocarbon dates.<br />

87<br />

<strong>Bova</strong> <strong>Marina</strong> <strong>Archaeological</strong> <strong>Project</strong> 1999<br />

Sample Material Trench Level ID Uncal. Calibrated Notes<br />

Date date*<br />

Umbro-1 Charcoal 3 210-220 Beta- 1770 +/- AD 120 - Late<br />

cm 122935 60 BP 415 (AD<br />

250)<br />

Roman<br />

Umbro-2 Charcoal 3 220-230 -- -- -- sample<br />

cm<br />

not<br />

analyzed<br />

Umbro-3 Charcoal 1 (-8n/- 181 cm, Beta- 4330 +/- 3045 - Copper<br />

37e) sublocus 1 122937 50 BP 2880 BC Age<br />

(Stratum<br />

(2910<br />

3)<br />

BC)<br />

Umbro-4 Charcoal 1 (-8n/- 210-220 Beta- 5790 +/- 4780 - Middle<br />

37e) cm (Base 122938 50 BP 4515 BC Neolithic<br />

<strong>of</strong> Stratum<br />

(4680<br />

3)<br />

BC)<br />

Umbro-5 Charcoal 1 (-8n/- 250-260 Beta- 6750 +/- 5685 - Early<br />

36e) cm 122939 50 BP 5565 BC Neolithic<br />

(Stratum<br />

(5600<br />

5)<br />

BC)<br />

Umbro-6 Animal 4 Stratum I Beta- modern modern AMS<br />

bone<br />

122940<br />

date; no<br />

valid<br />

result<br />

Umbro-7 Human 3 180-190 Beta- 1600 +/- AD 395- AMS<br />

bone<br />

cm 125061 40 BP 560 (AD date; Late<br />

440) Roman<br />

Umbro-13 Charcoal 1 (-7n/- 160-170 Beta- 5930 +/- 4945- AMS date<br />

36e) cm (Strata 135146 70 BP 4615 BC<br />

II/III)<br />

(4785<br />

BC)<br />

Umbro-15 Charcoal 1 (-7n/- 180-190 Beta- 6620 +/- 5660- AMS date<br />

36e) cm (top <strong>of</strong> 135147 60 BP 5485 BC<br />

Stratum<br />

(5610<br />

III)<br />

BC)<br />

Umbro-19 Charcoal 6 (-60n/- 60-70 (III: Beta- 3390 +/- 1865- AMS date<br />

6e) same level 135150 60 BP 1520 BC<br />

as base <strong>of</strong><br />

(1680<br />

pot scatter<br />

BC)<br />

* calibrated date is 2 sigma, 95% probability range with intersection <strong>of</strong> radiocarbon age with calibration curve in<br />

parentheses<br />

**Umbro-8, 9, 10, 11, 12, 14, 16, 17, 18, and 20 are charcoal samples archived and not dated.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!