Tropical forest conservation - unesdoc - Unesco [PDF]

va tropical, resalto el potencial del análisis de isótopos estables de carbono y de oxígeno de fósiles ...... This i

2 downloads 43 Views 37MB Size

Recommend Stories


patrimonio cinematográfico - unesdoc - Unesco [PDF]
Ficción. BN. Muda. Un estudiante pobre enamorado de una mujer rica vende su alma al diablo. DAS CABINET DES DR CALIGARI. (El gabinete del doctor Caligari). 19 19/20. aprox. 65 minutos, Producida por Decla Film. Dirigida por. Robert Wiene. Actores pr

Urban revitalization balanced between social ... - unesdoc - Unesco [PDF]
2004년 9월 유네스코가 바르셀로나에서. Following the round table organized. A la suite de la table ronde organisée by UNESCO's Social and Human Sciences par le Secteur des sciences sociales et 개최한 세계도시포럼전문가회의 이후 “역사. Sector, which took place during the UN- humaines

Forest (Conservation)
Pretending to not be afraid is as good as actually not being afraid. David Letterman

Forest conservation
Goodbyes are only for those who love with their eyes. Because for those who love with heart and soul

Tropical forest wood production
Before you speak, let your words pass through three gates: Is it true? Is it necessary? Is it kind?

tropical forest community ecology
Life isn't about getting and having, it's about giving and being. Kevin Kruse

Debt-for-Nature Initiatives and the Tropical Forest Conservation Act
You can never cross the ocean unless you have the courage to lose sight of the shore. Andrè Gide

Forest Conservation Ordinance
And you? When will you begin that long journey into yourself? Rumi

European Tropical Forest Research Network
If you feel beautiful, then you are. Even if you don't, you still are. Terri Guillemets

Tropical Food Forest Systems Workshop
Don’t grieve. Anything you lose comes round in another form. Rumi

Idea Transcript


TropicalForest Conservation Long-Term Processes of Human Evolution, Cultural Adaptations and Consumption Patterns

Long-Term Processes of Human Evolution, Cultural Adaptations and Consumption Patterns

Published in 2016 by the United Nations Educational, Scientific and Cultural Organization, 7, place de Fontenoy, 75352 Paris 07 SP, France and the UNESCO Office in Mexico, Presidente Masaryk 526, Polanco, Miguel Hidalgo, 11550 Ciudad de Mexico, D.F., Mexico.

© UNESCO 2016 ISBN: 978-92-3-000042-4

This publication is available in Open Access under the Attribution-ShareAlike 3.0 IGO (CC-BY-SA 3.0 IGO) license (http://creativecommons.org/ licenses/by-sa/3.0/igo/). By using the content of this publication, the users accept to be bound by the terms of use of the UNESCO Open Access Repository (http://www.unesco.org/open-access/terms-use-ccbysa-en).

The designations employed and the presentation of material throughout this publication do not imply the expression of any opinion whatsoever on the part of UNESCO concerning the legal status of any country, territory, city or area or of its authorities, or concerning the delimitation of its frontiers or boundaries.

The ideas and opinions expressed in this publication are those of the authors; they are not necessarily those of UNESCO and do not commit the Organization.

Original idea, concept, coordination and supervision of the editing and publication: The UNESCO Office in Mexico.

Conception, Edition and General Coordination of the project: Nuria Sanz, Head and Representative, UNESCO Office in Mexico Editorial work: Rachel Christina Lewis, UNESCO Office in Mexico José Pulido Mata, UNESCO Office in Mexico Chantal Connaughton, UNESCO Office in Mexico Graphic and cover design: Ananda Ramírez Cordero The UNESCO Office in Mexico would like to thank the Instituto de Ecología (INECOL) in Xalapa, the Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO) and the Gesellschaft für Internationale Zusammenarbeit (GIZ) GmbH for their collaboration. We would also like to extend our gratitude to all of the participants, whose contributions have made this publication possible.

Printed in Mexico.

Long-Term Processes of Human Evolution, Cultural Adaptations and Consumption Patterns

Table of contents Foreword  ǀ 6 Nuria Sanz 1. Tropical Rainforests as Long-Established Anthropomorphized Landscapes Tropical Rainforests as Long-Established Cultural Landscapes  ǀ 14 Robin Dennell Early Human Adaptation to Late Pleistocene-Holocene Rainforests in South Asia  ǀ 28 Patrick Roberts Sitios monumentales precolombinos en Amazonia   ǀ 44 Stéphen Rostain Xingu Garden Cities: Domesticated Forests of the Southern Amazon  ǀ 66 Michael Heckenberger Transformative Conservation in Three Dimensions: the LiDAR Recording of the Mosquitia Tropical Wilderness, Honduras  ǀ 88 Christopher T. Fisher 2. Interpreting the Past to Inform the Present and Implications for the Future: Lessons from Archaeology and Historical Ecology Morals to the Story of the ‘Mayacene’ from Geoarchaeology and Paleoecology  ǀ 110 Timothy Beach Un futuro sostenible para la Amazonia: lecciones de la arqueología  ǀ 140 José Iriarte Ancient Maya Water Management, Droughts and Urban Diaspora: Implications for the Present  ǀ 162 Lisa J. Lucero

Biodiversidad y resiliencia de la selva húmeda en Mesoamérica  ǀ 188 Sergio Guevara S.

3. Valuing Local and Indigenous Communities as Key Stakeholders in Forest Management and Conservation Valuing the Maya Forest as a Garden  ǀ 206 Anabel Ford A Tale of Three Species or the Ancient Soul of Tropical Forests   ǀ 228 Eduardo G. Neves The Failure of Systematic Conservation Planning and the Success of Community Conservation in the Sierra Norte of Oaxaca  ǀ 246 David Barton Bray Exploring Local Perspectives and Preferences in Forest Landscapes: Towards Democratic Conservation  ǀ 262 Douglas Sheil Valuing Traditional Knowledge for Conserving Biodiversity in Indonesia  ǀ 284 Herwasono Soedjito Perspectives from Local and Indigenous Producers in Mexico  ǀ 296 Various 4. Bridging the Research-Practice Gap Revisiting the Multifunctional Transition in Australia’s Wet Tropics: the Climate Change Crisis  ǀ 302 Stephen M. Turton Paisajes productivos sostenibles en México: de la idea a la realidad  ǀ 320 Martha Ileana Rosas Hernández 5. The Way Forward for More Sustainable Natural and Cultural Diversity All Over the World  ǀ 336 Nuria Sanz

Foreword Today we can consider the natural fragility of the green cloak of forest on the edge of the planet’s humid tropics a miracle. The surprise and admiration it evokes can be compared to the magnitude of its spatial scale and the sheer number of faunal and floral species it contains. More than a century ago, Euclides da Cuhna wrote that Amazonia was an unfinished chapter in the Book of Genesis. What is certain is

that today the United Nations 2030 Agenda for Sustainable Development obliges us to add more non-Amazonic pages to that narrative, because the forest corridors in Central America, Mexico, Central Africa and Island South-East Asian tropical archipelagos cannot be excluded from the story of Eden. All rainforests throughout the world have suffered from the forced displacements of populations, climate change, the devastating effects of natural disasters and the results of the growing predatory nature of humans. All forests have witnessed conflicts that debilitate societies and cultures. All of the world’s rainforests will soon face the pressures of human population growth and shifts. We are living in a century of mass migration and in the next two decades, a billion human beings are expected to change habitat. History is quickly moving us from one place to another, and those who arrive become the new societies who will have to connect to lands that are native to others. This is a familiar issue in Mexico, and from its experience very valid conclusions can be gained internationally: cultural identity and social respect for protected natural and cultural sites are safe-passages for conservation and are as important as regulations. Cultural programmes in protected natural areas should be ready to face the global challenge that migration will impose on conservation. The world is becoming increasingly smaller, but the rainforests must remain just as large.

6

Figure 1. UNESCO World Heritage site: Ancient Maya City and Protected Tropical Forests of Calakmul, Campeche, Mexico. © UNESCO Office in Mexico.

Rainforests and neighbourhoods are the livelihood of one billion of the world’s inhabitants and the need to reorient the ways to use them implies being able to regulate the carbon and water cycle, and the cycling of nutrients on Earth. The year 2011 was declared the International Year of Forests by the United Nations. In 2001, the World Heritage Committee adopted a specific policy for preserving forests. To date, 107 forest sites have been inscribed onto the UNESCO World Heritage List and cover a total of 75 million hectares across the globe. Fifty percent of the forests inscribed on the List are tropical and more than half are found in Latin America and the Caribbean. In this vast area, new forms of international cooperation must quickly be invented, as the challenges faced cannot be limited to the management capabilities of national environmental ministries. Annually, 13 million hectares of the world’s rainforests are lost. Illegal logging, slashing and burning, as well as the advance of the agricultural and cattle herding frontiers drive large discussions every year in the UNESCO Committees responsible for ensuring the global health of cultural and natural heritage. Forests, and especially tropical forests, are the most threatened natural spaces and therefore are very often represented on the UNESCO List of World Heritage in Danger. Governments, businesses, academia and civil society will have to multiply efforts in order to comply with their responsibilities of protecting rainforests to support the international community’s efforts. Recognizing the role of forests as carbon sinks

7

25

14

CAN

USA

USA

3

20 USA 28 USA 11

36

CAN

Chicago

5

USA

USA

USA

9

22 USA

46

47

USA

H

44

23

USA

USA

6

21 USA

Los Angeles

USA

USA

5

37 3

4

2

6

3

2

27

6

USA

ESP

MEX

41

16

TUN

1

MEX

MEX

20 MEX

MEX

MEX

24 MEX

Mexiko City

18 MEX

28

6

39

MEX

MEX

MEX

Pacific Ocean Océan Pacifique Océano Pacífico Pazifischer Ozean

21 MEX

7

26

29 MEX 32 40 MEX MEX 37 3 25 MEX MEX MEX 9 34 MEX

22

MEX

1

2

HTI

HTI

MEX

1

2

1

DOM

NER

16 26

35

USA

USA

1

USA

KNA

GTM

3

GTM

3

1

HND

1

NIC

2

4

3

3

1

2

SEN

2

2

3

BFA

GIN

NIC

CRI

F

MLI

3

1

GNB

2

CRI

SEN

1

SEN

SEN

COL

NIC

5

SEN

5

SLV

G

MRT

Caribbean Sea / Mer des Caraïbes Mar Caribe / Karibik

1

J

HND

2 3 HND

SLV

SLV

1

8

FRA

2

GTM

MEX

CUB

13

MEX

MEX

3

2

CUB

MEX

MEX

MEX

MEX

5

31 41

CUB

CUB

42

MEX

MEX

CUB

6

4

MEX

14

17

5

1

CUB

16

MEX

33

MEX

1

DZA

Havanna 30

12

38 36

ESP

USA

23 4

MEX

MEX

MEX

ESP

30

10

MEX

2

MEX

ESP

39

ESP

19 MEX

15

12

46

ESP

ESP

35

1

MAR

MAR

37

MEX

MEX

4

2

1

34 USA

3

Tunis

7

MAR

Houston

11 10

10

TUN

DZA

15

4

PRT

USA

USA

USA

3

5 4

TUN

MEX

USA

2

8

7

40

15

19

8

M

Rom 1

7 10

14

ITA

8 7

13

Casablanca

2

39

1

4

3

3 44

5

ESP

21 8

USA

USA

2

Madrid

34

25

P

31

38

USA

USA

1

18 22 23 36 27 20 9 29 42 32 19 24 28 26 31 47 35 17 8 33 11 45

1 10 48

PRT

USA

USA

6

3

PRT

24

USA

40 5

O PRT

2 9 PRT PRT 4

43

29

USA

USA

42

I

New York

4

3

COL

1

CRI

2

COL

GIN

TGO

Ibadan

2

1

CIV

GIN

3

GHA

Cali

NGA

Lagos

GHA

GHA

2

CMR

1

2

1

1

CIV

2

COL

Bogotá

1

CMR

BEN

1

2

2

VEN

PAN

Kano

2

1

BFA

GIN

1

PAN

1

NER BEN

1

4

Abidjan

1

CFA

3

CMR

1

1

1

VEN

COL

STP

1

1

COG

GAB

3

1

ECU

ECU

5

5

6

ECU

3

PER

2

ECU

Guayaquil

BRA

ECU

4

2

Fortaleza

ECU

2

5

BRA

PER

4

1

COG

Atlantic Ocean Océan Atlantique Océano Atlántico Atlantischer Ozean

1

COD

Kin

Luanda

BRA

PER

4

PER

Lima

2

PER

Salvador 3

2

BOL

1

BOL

BOL

3

BRA

5

CHL

6

2

BRA

PRY

5

11

1

ARG

BRA

ARG

8

1

10

RUS

Rio de Janeiro

São Paulo

PRY

ARG

9

2

ARG

6

ARG

is an area that still needs to be explored but at the same time is becoming a possible ARG

Figure 2. UNESCO Man and Biosphere (MaB) reserves in the Humid EST RUS Tropics. © UNESCO LVA Moscow FIN

SWE

25

1

RUS

1

9

16

ARG

RUS

33

34 26

1

5

Berlin

4

10

5

CZE

4

2

3 4 1 2 4

AUT 5 1 3 2

SVN

SVK 6

3

HUN 2

L

D

SRB MNE

Rom

1

2

1 1 1

ALB

5 4 10

15

4 3

16 3 13 4 11 9 6 14 5 8 1

N MKD

E

14

RUS

KAZ

ARG

3

33

14

CHN

RUS

21

11

RUS

MNG

RUS

2

15

19 6

RUS

MNG

22

MNG

ARG

CHN

4

4

RUS

CHL

MNG

KAZ

IRN

4

7

Mosul

s

1

3

2

Alexandria

1

1

2

JOR

1

8

Tehran

IRN

CHN

Baghdad

RUS

1

PKR

4

PKR

3

32

8

1

CHN

CHL

TKM

Taiyuan

2

IRN

Zhengzhou

Xi'an

18

3

JPN

JPN

1

JPN

6

Osaka

Busan

25

16 CHN

CHN

31 CHN

3

KOR

9

20

CHN

CHN

2

2

Yokohama Nagoya JPN

JPN

KOR

Shanghai

5

JPN

areas in the humid tropics. Its quality and scale areChongqing undisputed and the commitment CHN

Faisalabad 1 Delhi PAK

4

3

IND

CHN

26

24

Chengdu

8

Wuhan

12

CHN

4

CHN

JPN

Jaipurchallenge. must be equal to the These areas can become an area of best practices Kanpur Lucknow CHN

5

CHN 6 Xilin Gol (1987) CHN 7 Bogeda (1990) CHN 8 Shennongjia (1990) CHN 9 Yancheng (1992) CHN 10 Xishuangbanna (1993) CHN 11 Maolan (1996) CHN 12 Tianmushan (1996) CHN 13 Fenglin (1997) CHN 14 Jiuzhaigou Valley (1997) CHN 15 Nanji Islands (1998) CHN 16 Baishuijiang (2000) CHN 17 Gaoligong Mountain (2000) CHN 18 Huanglong (2000) CHN 19 Shankou Mangrove (2000) CHN 20 Baotianman (2001) CHN 21 Saihan Wula (2001) CHN 22 Dalai Lake (2002) CHN 23 Wudalianchi (2003) CHN 24 Yading (2003) CHN 25 Foping (2004) CHN 26 Qomolangma (2004) CHN 27 Chebaling (2007) CHN 28 Xingkai Lake (2007) CHN 29 Mao’er Mountain (2011) CHN 30 Jinggangshan (2012) CHN 31 Niubeiliang (2012) CHN 32 Snake Island - Laotie Mountain (2013) CHN 33 Hanma (2015) (la) Côte d’Ivoire (CIV) CIV 1 Taï (1977) CIV 2 Comoé (1983) Cameroon - le Cameroun - el Camerún - Kamerun (CMR) CMR 1 Waza (1979) CMR 2 Benoué (1981) CMR 3 Dja (1981) Democratic Republic of the Congo - la République démocratique du Congo - la República Democrática del Congo - Demokratische Republik Kongo (COD) COD 1 Luki (1976) COD 2 Yangambi (1976) COD 3 Lufira (1982) Congo - le Congo - le Congo - Kongo (COG) COG 1 Odzala (1977) COG 2 Dimonika (1988) Colombia - la Colombie - Colombia - Kolumbien (COL) COL 1 Cinturon Andino (1979) COL 2 El Tuparro (1979) COL 3 Sierra Nevada de Santa Marta (1979) COL 4 Ciénaga Grande de Santa Marta (2000) COL 5 Seaflower (2000) (le) Costa Rica (CRI) CRI 1 La Amistad (1982) CRI 2 Cordillera Volcánica Central (1988, ext. 2010) CRI 3 Agua y Paz (2007) Cuba - Cuba - Cuba - Kuba (CUB) CUB 1 Sierra del Rosario (1984) CUB 2 Baconao (1987) CUB 3 Cuchillas de Toa (1987) CUB 4 Península de Guanahacabibes (1987) CUB 5 Buenavista (2000) CUB 6 Ciénaga de Zapata (2000) Czech Republic - la République tchèque República Checa - Tschechische Republik (CZE) CZE 1 Krivoklátsko (1977) CZE 2 Trebon Basin(1977) CZE 3 Lower Morava (1986, ext.&ren. 2003)

7

KOR

KOR

KOR

CHN

PAK

JOR

1

4

PKR

5

2

IRN

EGY

2

Seoul Incheon Jinan have a universal responsibility in the Americas, but its responsibility should grow by Tokyo Daegu Maschhad

9

IRN

CHN

3

4

CHN

Changsha 30

IRN

IRN

8

RUS

fundamental. Amazonia, the Yucatán and the Meso-American Biological Corridor Pyongyang Tianjin Beijing

14

10 IRN 11

Basra

1

uch (2001) e - Chile (CHL) orge (1977, ext. 2012) ernández (1977) del Paine (1978) a San Rafael (1979) (1981) arias (1983, ext. 2010) mpana-Peñuelas (1984, ext. 2009) de Hornos (2005) es Templados Lluviosos de Los Australes (2007) dor Biológico Nevados de Chillán a de Laja (2011) hina - China (CHN) baishan (1979) ushan (1979) g (1979) gshan (1986) han (1987)

35

contagion and itsLahore efforts should be coordinated with the rest ofNanjing the protected natural

IRN

1

SYR

QAT

enata Sténa (1977) préné (1977) hina (1977) - Belarús - Weißrussland (BLR) nskiy (1978) zhskaya Pushcha (1993) olesie (2003; ext., ren., and transary with POL and UKR, 2012) Bolivia - Bolivien (BOL) Lajas (1977) la (1977) 1986) Brasil - Brasilien (BRA) Atlântica (including São Paulo City Belt) (1993, ext. 2002&2009) o (1994, ext. 2000&2001) nal (2000) ga (2001) l Amazon (2001) aço Range (2005) ublic - la République centrafricaine ana - Zentralafrikanische Republik (CAF) Lobaye (1977) gui-Bangoran (1979) - el Canadá - Kanada (CAN) Saint Hilaire (1978) on (1979) Point (1986) Mountain (1986) voix (1988) a Escarpment (1990) quot Sound (2000) int-Pierre (2000) Arrowsmith (2000) rry Lake (2000) West Nova (2001) and Islands - Frontenac Arch (2002) an Bay Littoral (2004) (2007) ouagan Uapishka (2007) ’Or Lake (2011) r Hills (2016) é (2016) sse - Suiza - Schweiz (CHE) stair - Parc Naziunal (1979, ext.&ren.

31

1

3

Kabul

6

ISR

ISR

Cairo

2

LBN

RUS

RUS

CHL

KGZ

IRN

IRN

LBN

LBN

4

CHN

32

PKR

1

60

2

1

KGZ

UZB

Izmir

1

28

CHN

1

1

CHN

6

CHN

6

TUR

Ankara

Harbin Changchun

Shenyang Our region’s role in the definition of sub regional conservation policies is

RUS

21

MNG

7

20

10

Istanbul

CHN

CHN

5

2

KAZ

1

1

13

23

MNG

6

RUS

RUS

Giza

22

KAZ

RUS

17

12 7

BGR

2

GRC

ARG

CHL

4

38

2

RUS

ARG

RUS

RUS

drivers of technical decision-making mechanisms.

RUS

2

7

27

RUS

10

39

12

9

1

UKR

RUS

coin’ and our task is to work collectively so that it is not funds that are the main

RUS

RUS

40 1

7

15

CHL

3

UKR

ARG

ARG

10

3

1

Cape Town

13

3

RUS

UKR

ROU

1

41

6

5

1

URY

emissions and curbing plant cover degradation operate as ‘two sides of the same 8

ARG

KAZ

8

7 4 5

RUS

Buenos Aires

4

ARG

12

RUS

UKR

2

RUS

CHL

23 7

K

2

URY

7

CHL

CHL

24

RUS

RUS

RUS

BLR

2

3

30

1

2

ITA 1

A

6 3 2 1 1

5 2

3

1

6

HRV

12

2 2 3 9 6

POL

5

37

RUS

29

1

3 8

B

1

3

LTU

5 7 12

28

RUS

8

CHL

RUS

13

3

1

financial market that needs to be better organized and discussed. Reducing carbon Santiago

36

RUS

RUS

1

RUS

1

RUS

St. Petersburg

2

4 2

CHN

5

IND

17

11

5

CHN

29

15 CHN

CHN

NewTaipei City

for cooperation between international Conventions, such as the Convention

CZE 4 Sumava (1990) CZE 5 Krkonoše (transboundary with POL, 1992) CZE 6 Bílé Karpaty (1996) Germany - l‘Allemagne - Alemania - Deutschland (DEU) DEU 1 Flusslandschaft Elbe (1979) DEU 2 Vessertal-Thüringer Wald (1979, ext. 1987&1990) DEU 3 Berchtesgadener Land (1990, ext.&ren. 2010) DEU 4 Schleswig-Holstenisches Wattenmeer, Halligen (1990, ext.&ren. 2004) 150 DEU 5 Schorfheide-Chorin (1990) DEU 6 Rhön (1991, ext. 2014) DEU 7 Spreewald (1991) DEU 8 Südost-Rügen (1991) DEU 9 Hamburgisches Wattenmeer (1992) DEU 10 Niedersächsisches Wattenmeer (1992) DEU 11 Vosges du Nord/Pfälzerwald (1992; transboundary with FRA, 1998) DEU 12 Oberlausitzer Heideund Teichlandschaft (1996) DEU 13 Schaalsee (2000) DEU 14 Bliesgau (2009) DEU 15 Schwäbische Alb (2009) Denmark - le Danemark - Dinamarca - Dänemark (DNK) DNK 1 North-East Greenland (1977) Dominican Republic - la République dominicaine República Dominicana - Dominikanische Republik (DOM) DOM 1 Jaragua-Bahoruco-Enriquillo (2002) Algeria - l‘Algérie - Argelia - Algerien (DZA) DZA 1 Tassili N’Ajjer (1986) DZA 2 El Kala (1990) DZA 3 Djurdjura (1997) DZA 4 Chrea (2002) DZA 5 Gouraya (2004) DZA 6 Taza (2004) DZA 7 Belezma (2015) DZA 8 Monts de Tlemcen (2016) Ecuador - l‘Équateur - el Ecuador - Ecuador (ECU) ECU 1 Archipiélago de Colón (Galápagos) (1984) ECU 2 Yasuní (1989) ECU 3 Sumaco (2000, ext. 2002) ECU 4 Podocarpus - El Condor (2007) ECU 5 Macizo del Cajas (2013) ECU 6 Bosque Seco (2014) Egypt - l‘Égypte - Egipto - Ägypten (EGY) EGY 1 Omayed (1981, ext. 1998) EGY 2 Wadi Allaqi (1993) Spain - l‘Espagne - España - Spanien (ESP) ESP 1 Grazalema (1977) ESP 2 Ordesa-Viñamala (1977, ext. 2013) ESP 3 Montseny (1978) ESP 4 Doñana (1980) ESP 5 La Mancha Húmeda (1980) ESP 6 La Palma (1983, ext.&ren. 1997&2002) ESP 7 Las Sierras de Cazorla y Segura (1983) ESP 8 Marismas del Odiel (1983) ESP 9 Urdaibai (1984) ESP 10 Sierra Nevada (1986) ESP 11 Cuenca Alta del Río Manzanares (1992) ESP 12 Lanzarote (1993) ESP 13 Menorca (1993, change in zonation 2004) ESP 14 Sierra de las Nieves y su Entorno (1995) ESP 15 Cabo de Gata-Nijar (1997) ESP 16 Isla de Hierro (2000) ESP 17 Bardenas Reales (2000)

ESP 18 ESP 19 ESP 20 ESP 21 ESP 22 ESP 23 ESP 25 ESP 24 ESP 26

Muniellos, Gran Cantábrica (2000, ext. 2003) Somiedo (2000) Redes (2001) Las Dehesas de Sierra Morena (2002) Terras do Miño (2002) Valle de Laciana, Gran Cantábrica (2003) Monfragüe (2003) Picos de Europa, Gran Cantábrica (2003) Valle de Jubera, Leza, Cidacos y Alhama (2003) ESP 27 120 Babia, Gran Cantábrica (2004) 90 ESP 28 Alto de Bernesga, Gran Cantábrica (2005) ESP 29 Área de Allariz (2005) ESP 30 Gran Canaria (2005) ESP 31 Los Argüellos, Gran Cantábrica (2005) ESP 32 Los Valles de Omaña y Luna (2005) ESP 33 Sierra del Rincón (2005) ESP 34 Las Sierras de Béjar y Francia (2006) ESP 35 Los Ancares Leoneses, Gran Cantábrica (2006) ESP 36 Los Ancares Lucenses y Montes de Cervantes, Navia y Becerrea, Gran Cantábrica (2006) ESP 37 Reserva de la Biosfera intercontinental del Mediterraneo (transb. with MAR, 2006) ESP 38 Rio Eo, Oscos y Terras de Buron (2007) ESP 39 Fuerteventura (2009) ESP 40 Gerês (transboundary with PRT, 2009) ESP 41 La Gomera (2012) ESP 42 Las Ubinas - La Mesa (2012) ESP 43 Marinas Corunesas e Terras do Mandeo (2013) ESP 44 Terres de l‘Ebre (2013) ESP 45 Real Sitio de San Ildefonso - El Espinar (2013) ESP 46 Macizo de Anaga (2015) ESP 47 Meseta Ibérica (transboundary with PRT, 2015) ESP 48 Tejo/Tajo Internacional (transboundary with PRT, 2016) Estonia - l‘Estonie - Estonia - Estland (EST) EST 1 West-Estonian Archipelago (1990) Ethiopia - l‘Éthiopie - Etiopía - Äthiopien (ETH) ETH 1 Kafa (2010) ETH 2 Yayu (2010) ETH 3 Sheka (2012) ETH 4 Lake Tana (2015) Finland - la Finlande - Finlandia - Finnland (FIN) FIN 1 North Karelian (1992) FIN 2 Achipelago Sea Area (1994) France - la France - Francia - Frankreich (FRA) FRA 1 Camargue (Rhône-Delta,1977, ext.&ren. 2006) FRA 2 Commune de Fakarava (1977, ext.&ren. 2006) FRA 3 Vallée du Fango (1977, ext. 1990) FRA 4 Cévennes (1984) FRA 5 Iles et Mer d‘Iroise (1988, ext.&ren. 2012) FRA 6 Vosges du Nord / Pfälzerwald (1988; transboundary with DEU, 1998) FRA 7 Mont Ventoux (1990) FRA 8 Archipel de la Guadeloupe (1992) FRA 9 Luberon-Lure (1997, ext.&ren. 2010) FRA 10 Fontainebleau et du Gâtinais (1998, ext.&ren. 2010) FRA 11 Bassin de la Dordogne (2012) FRA 12 Marais Audomarois (2013) FRA 13 Mont-Viso (transboundary with ITA, 2013) FRA 14 Gorges du Gardon (2015)

CHN

CHN

Micronesia - la Micronésie - Micronesia - Mikronesien (FSM) FSM 1 Utwe (2005) FSM 2 And Atoll (2007) Gabon - le Gabon - el Gabón - Gabun (GAB) GAB 1 Ipassa-Makokou (1983) United Kingdom - le Royaume-Uni - el Reino Unido Vereinigtes Königreich (GBR) GBR 1 Wester Ross (1976, ext. and ren. 2016) GBR 2 Braunton Burrows - North Devon (1976, ext. 2002) GBR 3 Biosffer Dyfi60 (1976, ext.&ren. 2009) GBR 4 Galloway and Southern Ayrshire (2012) GBR 5 Brighton and Lewes Downs (2014) GBR 6 Isle of Man (2016) (le) Ghana (GHA) GHA 1 Bia (1983) GHA 2 Songor (2011) GHA 3 Lake Bosomtwe (2016) Guinea - la Guinée - Guinea - Guinea (GIN) GIN 1 Massif du Ziama (1980) GIN 2 Monts Nimba (1980) GIN 3 Badiar (2002) GIN 4 Haut Niger (2002) (la) Guinea-Bissau (GNB) GNB 1 Boloma Bijagós (1996) Greece - la Grèce - Grecia - Griechenland (GRC) GRC 1 Gorge of Samaria (1981) GRC 2 Mount Olympus (1981) (le) Guatemala (GTM) GTM 1 Maya (1990) GTM 2 Sierra de Las Minas (1992) GTM 3 Trifinio Fraternidad (transboundary with SLV and HND, 2011, ext. 2016) (le) Honduras (HND) HND 1 Río Plátano (1980) HND 2 Trifinio Fraternidad (transboundary with SLV and GTM, 2011, ext. 2016) HND 3 Cacique Lempira, Señor de las Montañas (2015) Croatia - la Croatie - Croacia - Kroatien (HRV) HRV 1 Velebit Mountain (1977) HRV 2 Mura Drava Danube (transboundary with HUN, 2012) Haiti (HTI) HTI 1 La Selle (2012) HTI 2 La Hotte (2016) Hungary - la Hongrie - Hungría - Ungarn (HUN) HUN 1 Aggtelek (1979) HUN 2 Hortobágy (1979) HUN 3 Kiskunság (1979) HUN 4 Lake Fertö (1979) HUN 5 Pilis (1980) HUN 6 Mura Drava Danube (transboundary with HRV, 2012) Indonesia - l‘Indonésie - Indonesia - Indonesien (IDN) IDN 1 Cibodas (1977) IDN 2 Komodo (1977) IDN 3 Lore Lindu (1977) IDN 4 Tanjung Puting (1977) IDN 5 Gunung Leuser (1981) IDN 6 Siberut (1981) IDN 7 Giam Siak Kecil - Bukit Batu (2009) IDN 8 Wakatobi (2012) IDN 9 Bromo Tengger Semeru-Arjuno (2015) IDN 10 Taka Bonerate-Kepulauan Selayar (2015) IDN 11 Balambangan (2016)

India - l‘Inde - la India - Indien (IND) IND 1 Nilgiri (2000) IND 2 Gulf of Mannar (2001) IND 3 Sunderban (2001) IND 4 Nanda Devi (2004) IND 5 Nokrek (2009) IND 6 Pachmarhi (2009) IND 7 Similipal (2009) IND 8 Achanakmar-Amarkantak (2012) IND 9 Great Nicobar (2013) 0 IND30 10 Agasthyamala (2016) Ireland - l‘Irlande - Irlanda - Irland (IRL) IRL 1 Dublin Bay (1981, renamed 2015) IRL 2 Killarney (1982) Iran - Iran - Irán - Iran (IRN) IRN 1 Arasbaran (1976) IRN 2 Arjan (1976) IRN 3 Geno (1976) IRN 4 Golestan (1976) IRN 5 Hara (1976) IRN 6 Kavir (1976) IRN 7 Lake Oromeeh (1976) IRN 8 Miankaleh (1976) IRN 9 Touran (1976) IRN 10 Dena (2010) IRN 11 Tang-e-Sayad and Sabzkuh (2015) IRN 12 Hamoun (2016) Israel - Israël - Israel - Israel (ISR) ISR 1 Mount Carmel (1996) ISR 2 Ramat Menashe (2011) Italy - l‘Italie - Italia - Italien (ITA) ITA 1 Circeo (1977) ITA 2 Collemeluccio-Montedimezzo (1977) ITA 3 Miramare (1979) ITA 4 Cilento and Valle di Diano (1997) ITA 5 Somma-Vesuvio and Miglio d’Oro (1997) ITA 6 Valle del Ticino (2002) ITA 7 Tuscan Islands (2003) ITA 8 Toscana (2004, ext. and ren. 2016) ITA 9 Area della Biosfera del Monviso (transboundary with FRA, 2013) ITA 10 Sila (2014) ITA 11 Ledro Alps and Judicaria (2015) ITA 12 Po Delta (2015) ITA 13 Appennino Tosco-Emiliano (2015) ITA 14 Collina Po (2016) Jordan - la Jordanie - Jordania - Jordanien (JOR) JOR 1 Dana (1998) JOR 2 Mujib (2011) Japan - le Japon - el Japón - Japan (JPN) JPN 1 Mount Hakusan (1980, ext. 2016) JPN 2 Mount Odaigahara, Mount Omine and Osugidani (1980, ext. and ren. 2016) JPN 3 Shiga Highland (1980, ext. 2014) JPN 4 Yakushima and Kuchinoerabu Jima (1980, ext. and ren. 2016) JPN 5 Aya (2012) JPN 6 Minami Alps (2014) JPN 7 Tadami (2014) Kazakhstan - le Kazakhstan - Kazajstán - Kasachstan (KAZ) KAZ 1 Korgalzhyn (2012) KAZ 2 Alakol (2013) KAZ 3 Ak-Zhayik (2014) KAZ 4 Katon-Karagay (2014) KAZ 5 Aksu-Zhabagly (2015) KAZ 6 Barsakelmes (2016)

on Biological Diversity, the Ramsar Convention, the UNESCO Man and the Biosphere (MAB) Programme and the UNESCO World Heritage Convention.

Cooperation between international Conventions is a recommended but never fully achieved task. Furthermore, safeguarding forms of regional connectivity is a pending and pressing task. It requires, in addition to technical platforms, finding

a space in dialogue at scientific and industrial summits and political cooperation scenarios. In this way, interaction with traditional productive practices is essential and urgent, and this belief justifies the aim of our efforts.

Rainforests are also areas where greater efforts are needed in terms of applied

research for conversation. Science must be understood above all as anthropological and sociological work with the human communities who live within the rainforests

or on their periphery. Forms of sustainable development are always designed based on consultations or participative work. Over the last 40 years, development

anthropology or anthropology for development have not stopped clamouring for a methodology that starts with the cultural understanding of expectations.

Kenya - le Ke KEN 1 KEN 2 KEN 3 KEN 4 KEN 5 KEN 6 Kyrgyzstan - l KGZ 1 KGZ 2 Cambodia - le KHM 1 Saint Kitts an Saint Kitts y KNA 1 Republic of K República de KOR 1 KOR 2 KOR 3 KOR 4 KOR 5 Lebanon - le LBN 1 LBN 2 LBN 3 Sri Lanka (LK LKA 1 LKA 2 LKA 3 LKA 4 Lithuania - la LTU 1 Latvia - la Let LVA 1 Morocco - le MAR 1 MAR 2 MAR 3

MAR 4 Madagascar MDG 1 MDG 2 MDG 3 MDG 4 Maldives - les MDV 1 Mexico - le M MEX 1 MEX 2 MEX 3 MEX 4 MEX 5 MEX 6 MEX 7 MEX 8 MEX 9 MEX 10 MEX 11 MEX 12 MEX 13 MEX 14 MEX 15 MEX 16

MNE

1

1

1 1

ALB

16 13 4

2

RUS

7

N 11 14 3 9 BGR 5 6 8 MKD 1 GRC

RUS

6

CHN

5

10

1

4

7

Mosul

2

Alexandria

1

1

2

Baghdad

SYR

LBN

2

12 IRN

2

PAK

2

IRN

CHN

18

5

Jaipur

IRN

IRN

1

QAT

Ahmedabad

Mumbai Khartum

2

2

2

3

CHN

ETH

Seoul 1 Daegu Osaka Busan 3 5

17

IND

3

IND

MRR

3

Guangzhou 2 Shenzhen 27 Hong Kong

CHN

Tropic of cancer / Tropique du C Trópico de Cáncer / Nördlicher W

CHN

VNM

6

2

THA

19

3

4

VNM

VNM

THA

Pacific Ocean Océan Pacifique Océano Pacífico Pazifischer Ozean

Yangon Chennai

VNM

Bangkok

2

1

4

2

5

THA

IND

3

PHL

PHL

KHM

2

1

9

VNM

Ho Chi Minh City

1

PHL

VNM

VNM

1

IND

Manila Quezon City

7

VNM

10

CHN

NewTaipei City

CHN

1

IND

1

29

CHN

CHN

10

1

YEM

4

JPN

15

CHN

Chittagong

5

JPN

Shanghai

5

Tokyo 6 Yokohama Nagoya JPN

2

JPN

12

CHN

JPN

KOR

CHN

11

CHN

2

9

3

JPN

JPN

CHN

Changsha

7

KOR

KOR

Nanjing Wuhan

CHN

7

IND

Nagpur

Addis Abeba

SDN

8

VNM

LKA

ETH

ETH

8

4

THA

1

4

1

4

KOR

30

5

Pune Hyderabad

ETH

2

8

Bengaluru

1

SDN

4

PKR

Incheon

CHN

1

Omdurman

CFA

Chengdu

CHN

Sanaa

YEM

PKR

Kolkata

IND

Surat

CHN

Chongqing

24

Kanpur Lucknow

6

Jeddah

2

EGY

20

CHN

CHN

26

IND

Karachi

1

ARE

CHN

3

IND

PAK

CHN

Riyadh

2

Zhengzhou

31

CHN

16

CHN

4

Faisalabad 1 Delhi

EGY

3

CHN

Jinan

Xi'an 25

14

Lahore

IRN

Basra

JOR

32

Pyongyang

KOR

Kabul

10 IRN 11

ISR

1

Taiyuan

Maschhad

IRN

IRN

JOR

1

9

IRN

6

LBN

ISR

Cairo Giza

8

Tehran

3

1

TKM

IRN

IRN

RUS

1

PKR

3

Tianjin

1

IRN

LBN

Beijing

PKR

Izmir

1

KGZ

35 31 RUS

1

CHN

2

1

KGZ

UZB

TUR

Ankara

2

LKA

3

4

1

9

PLW

IND

LKA

2

2

LKA

FSM

MYS

MDV

FA

2

2

2

1

1

IDN

2 KEN KEN

COD

5

MYS

7

6

UGA

Singapur

1

3

KEN

1

Nairobi

2

RWA

Equator / Ecuador te

IDN

KEN

UGA

5

TZA

1

4

IDN

6

KEN

IDN

4

3

KEN

1

FSM

1

1

IDN

KEN

TZA

3

nshasa

TZA

8

4

IDN

Jakarta

TZA

10

1

IDN

9 Surabaya 11

IDN

2

IDN

IDN

a

n

Shenyang

CHN

KAZ

1

1

Istanbul

2

IDN

Indian Ocean Océan Indien Océano Índico Indischer Ozean

3

COD

2

MDG

2

MWI

1

1

1

4

4

MUS

3

ZAF

ZAF

AUS

MDG

MDG

6

3

1

MWI

ZWE

3

MDG

ZAF

8

14

5

ZAF

Johannesburg

Tropic of capricorn / Tropique du C Trópico de Capricornio / Südlicher W

AUS

13

AUS

AUS

It’s time that social and human disciplines become the subject of research. Social 6

AUS

Durban

and natural scientists need to plan together the way in which to foresee the

60

7

AUS

2

5

ZAF

ZAF

7

8

ZAF

enya - Kenya - Kenia (KEN) Mount Kenya (1978) Mount Kulal (1978) Malindi-Watamu (1979) Kiunga (1980) Amboseli (1991) Mount Elgon (2003) le Kirghizistan - Kirguistán - Kirgisistan (KGZ) Sary-Chelek (1978) Issyk Kul (2001) 30 - Camboya - Kambodscha (KHM) e Cambodge Tonle Sap (1997) nd Nevis - Saint-Kitts-et-Nevis Nevis - St. Kitts und Nevis (KNA) St. Mary‘s (2011) Korea - République de Corée e Corea - Republik Korea (KOR) Mount Sorak (1982, ext. 2016) Jeju Island (2002) Shinan Dadohae (2009, ext. 2016) Gwangneung Forest (2010) Gochang (2013) Liban - el Libano - Libanon (LBN) Shouf (2005) Jabal Al Rihane (2007) Jabal Moussa (2009) KA) Hurulu (1977) Sinharaja (1978) Kanneliya-Dediyagala-Nakiyadeniya (2004) Bundala (2005) a Lituanie - Lituania - Litauen (LTU) Zuvintas (2011) ttonie - Letonia - Lettland (LVA) North Vidzeme (1997) Maroc - Marruecos - Marokko (MAR) Arganeraie (1998) Oasis du sud marocain (2000) Réserve de biosphère intercontinentale de la Méditerranée (transboundary with ESP, 2006) Atlas Cedar (2016) - Madagaskar (MDG) Mananara Nord (1990) Sahamalaza-Iles Radama (2001) Littoral de Toliara (2003) Belo-sur-Mer—Kirindy-Mitea (2016) s Maldives - Maldivas - Malediven (MDV) Baa Atoll (2011) Mexique - México - Mexiko (MEX) Mapimí (1977) La Michilía (1977) Montes Azules (1979) El Cielo (1986) Sian Ka’an (1986) Sierra de Manantlán (1988) Région de Calakmul (1993, ext.&ren. 2006) Alto Golfo de California (1993, ext. 1995) El Triunfo (1993) El Vizcaíno (1993) Islas de Golfo de California (1995) Sierra Gorda (2001) Banco Chinchorro (2003) Ría Celestún (2003) Sierra La Laguna (2003) Ría Lagartos (2004)

Sydney

4

sustainable development of rural areas and of their adaptive capacity to changes. AUS

1

AUS

12

ZAF

AUS

2

9

AUS

Melbourne

AUS

1

The world’s rainforests have not been lifeless territories. For thousands of years, 11 AUS

AUS

10 AUS

they have contained innumerable forms of cultural adaptations along the world’s latitudes. The task ahead is to know how to read, without wasting time or losing any details, all of the social responses and to be able to learn from the rainforest’s long-term cultural history. 60

The planet’s enormous tropical, green cloak immediately takes us back to an 30

original and primitive world, even if it has been and cultivated over Polar circle managed / Cercle polaire Círculo polar / Polarkreis

millennia. We must deal with it from the belief of the advantages of preservation MEX 17 Barranca de Metztilán (2006) MEX 18 Chamela-Cuixmala (2006) MEX 19 Cuatro Ciénagas (2006) MEX 20 Cumbres de Monterrey (2006) MEX 21 Huatulco (2006) MEX 22 La Encrucijada (2006) MEX 23 Laguna Madre y Delta de Río Bravo (2006) MEX 24 La Primavera (2006) MEX 25 La Sepultura (2006) MEX 26 Los Tuxtlas (2006) 90 60 MEX 27 Maderas del Carmen, Coahuila (2006) MEX 28 Mariposa Monarca (2006) MEX 29 Pantanos de Centla (2006) MEX 30 Arrecife Alacranes (2006) MEX 31 Sistema Arrecifal Veracruzano (2006) MEX 32 Selva El Ocote (2006) MEX 33 Sierra de Huautla (2006) MEX 34 Volcan Tacaná (2006) MEX 35 Sierra de Alamos - Rio Cuchujaqui (2007) MEX 36 Islas Marietas (2008) MEX 37 Lagunas de Montebello (2009) MEX 38 Islas Marías (2010) MEX 39 Los Volcanes (2010) MEX 40 Nahá-Metzabok (2011) MEX 41 Tehuacán-Cuicatlán (2012) MEX 42 Isla Cozumel (2016) Former Yugoslav Republic of Macedonia - l‘ex-République yougoslave de Macédoine - ex República Yugoslava de Macedonia - ehemalige jugoslawische Republik Mazedonien (MKD) MKD 1 Ohrid - Prespa (transboundary with ALB, 2014) (le) Mali (MLI) MLI 1 Boucle du Baoulé (1982) (le) Myanmar (MMR) MMR 1 Inlay Lake (2015) Montenegro - le Monténégro - Montenegro Montenegro (MNE) MNE 1 Tara River Basin (1976) Mongolia - la Mongolie - Mongolia - Mongolei (MNG) MNG 1 Great Gobi (1990) MNG 2 Boghd Khan Uul (1996) MNG 3 Uvs Nuur Basin (1997) MNG 4 Hustai Nuruu (2002) MNG 5 Dornod Mongol (2005) MNG 6 Mongol Daguur (2007) Mauritania - la Mauritanie - Mauritania - Mauretanien (MRT) MRT 1 Delta du Fleuve Sénégal (transboundary with SEN, 2005) Mauritius - Maurice - Mauricio - Mauritius (MUS) MUS 1 Macchabee / Bel Ombre (1977) (le) Malawi (MWI) MWI 1 Mount Mulanje (2000) MWI 2 Lake Chilwa Wetland (2006) Malaysia - la Malaisie - Malasia - Malaysia (MYS) MYS 1 Tasik Chini (2009) MYS 2 Crocker Range (2014) Niger - le Niger - el Níger - Niger (NER) NER 1 W Region (1996; ext.& transboundary with BEN and BFA, 2002) NER 2 Aïr et Ténéré (1997) Nigeria - le Nigéria - Nigeria - Nigeria (NGA) NGA 1 Omo (1977) (le) Nicaragua (NIC) NIC 1 Bosawas (1997) NIC 2 Río San Juan (2003) NIC 3 Ometepe Island (2010)

and not from a nostalgic point of view. The magnitude of this green cloak obliges

Netherlands - les Pays-Bas - los Países Bajos Niederlande (NLD) NLD 1 Wadden Sea Area (1986) Pakistan - le Pakistan - el Pakistán - Pakistan (PAK) PAK 1 Lal Suhanra (1977) PAK 2 Ziarat Juniper Forest (2013) Panama - le Panama - Panamá - Panama (PAN) PAN 1 Darién (1983) PAN 2 La Amistad (2000) Peru - le Pérou - el Perú - Peru (PER) PER 1 Huascarán (1977) 150 120 PER 2 Manu (1977) PER 3 Noroeste Amotapes – Manglares (1977, ext. and ren. 2016) PER 4 Oxapampa-Ashaninka-Yanesha (2010) PER 5 Gran Pajatén (2016) Philippines - les Philippines - Filipinas - Philippinen (PHL) PHL 1 Palawan (1977) PHL 2 Puerto Galera (1977) PHL 3 Albay (2016) Palau - Palaos - Palau - Palau (PLW) PLW 1 Ngaremeduu (2005) Poland - la Pologne - Polonia - Polen (POL) POL 1 Babia Gora (1976, ext. 1997&2001) POL 2 Bialowieza (1976, ext. 2005) POL 3 Lukajno Lake (1976) POL 4 Slowinski (1976) POL 5 Karkonosze (transboundary with CZE, 1992) POL 6 Tatra (transboundary with SVK, 1992) POL 7 East Carpathians (transboundary with SVK and UKR, 1998) POL 8 Puszcza Kampinoska (2000) POL 9 West Polesie (2002; ext., ren.&transboundary with UKR and BLR, 2012) POL 10 Tuchola Forest (2010) Democratic People‘s Republic of Korea la République populaire démocratique de Corée República Popular Democrática de Corea Demokratische Volksrepublik Korea (PRK) PRK 1 Mount Paekdu (1989) PRK 2 Mount Kuwol (2004) PRK 3 Mount Myohyang (2009) PRK 4 Mount Chilbo (2014) (le) Portugal (PRT) PRT 1 Paúl do Boquilobo (1981) PRT 2 Corvo Island (2007) PRT 3 Graciosa Island (2007) PRT 4 Flores Island (2009) PRT 5 Xurés (transboundary with ESP, 2009) PRT 6 Berlengas (2011) PRT 7 Santana Madeira (2011) PRT 8 Meseta Ibérica (transb. with ESP, 2015) PRT 9 Fajãs de São Jorge (2016) PRT 10 Tejo/Tajo Internacional (transb. ESP, 2016) Paraguay - le Paraguay - el Paraguay - Paraguay (PRY) PRY 1 Bosque Mbaracayú (2000) PRY 2 El Chaco (2005) Qatar - le Qatar - Qatar - Katar (QAT) QAT 1 Al-Reem (2007) Romania - la Roumanie - Rumania - Rumänien (ROU) ROU 1 Pietrosul Mare (1979) ROU 2 Retezat (1979) ROU 3 Danube Delta (1992; tbr. with UKR, 1998) Russian Federation - Fédération de Russie Federación de Rusia - Russische Föderation (RUS) RUS 1 Kavkazskiy (1978)

RUS 2 Okskiy (1978, pt. of Oka until 2000) RUS 3 Prioksko-Terrasnyi (1978, pt. of Oka until 2000) RUS 4 Sikhote-Alin (1978) RUS 5 Tsentral’nochernozem (1978) RUS 6 Astrakhanskiy (1984) RUS 7 Kronotskiy (1984) RUS 8 Laplandskiy (1984) RUS 9 Pechoro-Ilychskiy (1984) RUS 10 Sayano-Shushenskiy (1984) RUS 11 Sokhondinskiy (1984) RUS 12 Voronezhskiy (1984) 180 RUS 13 Tsentralnolesnoy (1985) RUS 14 Baikalskyi (1986, pt of Lake Baikal until 2000) RUS 15 Barguzinskyi (1986, pt of Lake Baikal until 2000) RUS 16 Tsentralnosibirskiy (1986) RUS 17 Chernyje Zemli (1993) RUS 18 Taimyrsky (1995) RUS 19 Daursky (1997) RUS 20 Teberda (1997) RUS 21 Ubsunorskaya Kotlovina (1997) RUS 22 Katunskiy (2000) RUS 23 Nerusso-Desnianskoe-Polesie (2001) RUS 24 Visimskiy (2001) RUS 25 Vodlozersky (2001) RUS 26 Darvinskiy (2002) RUS 27 Commander Islands (2002) RUS 28 Nijegorodskoe Zavolje (2002) RUS 29 Smolensk Lakeland (2002) RUS 30 Ugra (2002) RUS 31 Far East Marine (2003) RUS 32 Kedrovaya Pad (2004) RUS 33 Kenozersky (2004) RUS 34 Valdaiskiy (2004) RUS 35 Khankaiskiy (2005) RUS 36 Middle Volga Integrated Biosphere (2006) RUS 37 Great Volzhsko-Kamsky (2007) RUS 38 Rostovsky (2008) RUS 39 Altaisky (2009) RUS 40 Wolga-Akhtuba Floodplain (2011) RUS 41 Bashkirskiy Ural (2012) Rwanda - le Rwanda - Rwanda - Ruanda (RWA) RWA 1 Volcans (1983) Sudan - le Soudan - el Sudán - Sudan (SDN) SDN 1 Dinder (1979) SDN 2 Radom (1979) Senegal - le Sénégal - el Senegal - Senegal (SEN) SEN 1 Samba Dia (1979) SEN 2 Delta du Saloum (1980) SEN 3 Niokolo-Koba (1981) SEN 4 Delta du Fleuve Sénégal (transboundary with MRT, 2005) SEN 5 Ferlo (2012) El Salvador (SLV) SLV 1 Apaneca - Llamatepec (2007) SLV 2 Xiriualtique - Jiquitizco (2007) SLV 3 Trifinio Fraternidad (transboundary with GTM and HND, 2011, ext. 2016) Serbia - la Serbie - Serbia - Serbien (SRB) SRB 1 Golija-Studenica (2001) Sao Tome and Principe - Sao Tomé-et-Principe Santo Tomé y Príncipe - São Tomé und Príncipe (STP) STP 1 The Island of Príncipe (2012) Slovakia - la Slovaquie - Eslovaquia - Slowakei (SVK) SVK 1 Slovenskiý Kras (1977) SVK 2 Polana (1990) SVK 3 Tatra (transboundary with POL, 1992)

SVK 4

East Carpathians (transboundary with POL and UKR, 1998) Slovenia - la Slovénie - Eslovenia - Slowenien (SVN) SVN 1 Julian Alps (2003) SVN 2 The Karst (2004) SVN 3 Kozjansko and Obsotelje (2010) Sweden - la Suède - Suecia - Schweden (SWE) SWE 1 Kristianstad Vattenrike (2005) SWE 2 Lake Vänern Archipelago (2010) SWE 3 Blekinge Archipelago (2011) SWE 4 Nedre Dalälven River Landscape (2011) SWE 5 East Vättern Scarp Landscape (2012) Syria - la Syrie - Siria - Syrien (SYR) SYR 1 Lajat (2009) Togo - le Togo - el Togo - Togo (TGO) TGO 1 Complexe Oti-Keran / Oti-Mandouri (2011) Thailand - la Thaïlande - Tailandia - Thailand (THA) THA 1 Sakaerat (1976) THA 2 Hauy Tak Teak (1977) THA 3 Mae Sa-Kog Ma (1977) THA 4 Ranong (1997) Turkmenistan - le Turkménistan - Turkmenistan Turkmenistan (TKM) TKM 1 Repetek (1978) Tunisia - la Tunisie - Túnez - Tunesien (TUN) TUN 1 Djebel Bou-Hedma (1977) TUN 2 Djebel Chambi (1977) TUN 3 Ichkeul (1977) TUN 4 Iles Zembra et Zembretta (1977) Turkey - la Turquie - Turquía - Türkei (TUR) TUR 1 Camili (2005) Tanzania - la Tanzanie - Tanzanía - Tansania (TZA) TZA 1 Lake Manyara (1981) TZA 2 Serengeti-Ngorongoro (1981) TZA 3 East Usambara (2000) TZA 4 Jozani-Chwaka Bay (2016) Uganda - l‘Ouganda - Uganda - Uganda (UGA) UGA 1 Queen Elizabeth (1979) UGA 2 Mount Elgon (2005) Ukraine - l‘Ukraine - Ucrania - Ukraine (UKR) UKR 1 Chernomorskiy (1985) UKR 2 Askaniya-Nova (1985) UKR 3 Carpathian (1992) UKR 4 Danubel Delta (transboundary with ROU, 1998) UKR 5 East Carpathians (transboundary with POL and SVK, 1998) UKR 6 West Polesie (2002; ext., ren.&transboundary with POL and BLR, 2012) UKR 7 Desnianskyi (2009) UKR 8 Roztochya (2011) Uruguay - l‘Uruguay - el Uruguay - Uruguay (URY) URY 1 Bañados del Este (1976) URY 2 Bioma Pampa-Quebradas del Norte (2014) United States - les États-Unis - los Estados Unidos Vereinigte Staaten (USA) USA 1 Aleutian Islands (1976) USA 2 Beaver Creek (1976) USA 3 Big Bend (1976) USA 4 Cascade Head (1976) USA 5 Central Plains (1976) USA 6 Channel Islands (1976) USA 7 Coram (1976) USA 8 Denali (1976) USA 9 Desert (1976) USA 10 Everglades (1976)

us to think of another way of cohabiting.

A Contribution from Mexico

USA 11 Fraser (1976) USA 12 Glacier (1976) USA 13 H.J. Andrews (1976) USA 14 Hubbard Brook (1976) USA 15 Jornada (1976) USA 16 Luquillo (1976) USA 17 Noatak (1976) USA 18 Olympic (1976) USA 19 Organ Pipe Cactus (1976) USA 20 Rocky Mountain (1976) USA 21 San Dimas (1976) USA 22 San Joaquin (1976) USA 23 Sequoia-Kings Canyon (1976) USA 24 Stanislaus-Tuolumne (1976) USA 25 Three Sisters (1976) USA 26 Virgin Islands (1976) USA 27 Yellowstone (1976) USA 28 Konza Prairie (1978) USA 29 University of Michigan Biological Station (1979) USA 30 Niwot Ridge (1979) USA 31 Virginia Coast (1979) USA 32 Hawaiian Islands (1980) USA 33 Isle Royale (1980) USA 34 Big Thicket (1981) USA 35 Guanica (1981) USA 36 California Coast Ranges (1983) USA 37 Central Gulf Coast Plain (1983) USA 38 South Atlantic Coastal Plain (1983) USA 39 Mojave and Colorado Deserts (1984) USA 40 Carolinian-South Atlantic (1986) USA 41 Glacier Bay-Admiralty Islands (1986) USA 42 Golden Gate (1986) USA 43 New Jersey Pinelands (1988) USA 44 Southern Appalachian (1988) USA 45 Champlain-Adirondak (1989) USA 46 Mammoth Cave Area (1990, ext. 1996) USA 47 Land Between the Lakes Area (1991) Uzbekistan - l‘Ouzbékistan - Uzbekistán - Usbekistan (UZB) UZB 1 Mount Chatkal (1978) Venezuela - Venezuela - Venezuela - Venezuela (VEN) VEN 1 Alto Orinoco-Casiquiare (1993) VEN 2 Delta Orinoco (2009) Viet Nam - le Viet Nam - Viet Nam - Vietnam (VNM) VNM 1 Can Gio Mangrove (2000) VNM 2 Dong Nai (2001, ext.&ren. 2011) VNM 3 Cat Ba (2004) VNM 4 Red River Delta (2004) VNM 5 Kien Giang (2006) VNM 6 Western Nghe An (2007) VNM 7 Cu Lao Cham - Hoi An (2009) VNM 8 Mui Ca Mau (2009) VNM 9 Langbiang (2015) Yemen - le Yémen - el Yemen - Jemen (YEM) YEM 1 Socotra Archipelago (2003) YEM 2 Bura’a (2011) South Africa - l‘Afrique du Sud - Sudáfrica - Südafrika (ZAF) ZAF 1 Kogelberg (1998) ZAF 2 Cape West Coast (2000, ext. 2003) ZAF 3 Kruger To Canyons (2001) ZAF 4 Waterberg (2001) ZAF 5 Cape Winelands (2007) ZAF 6 Vhembe (2009) ZAF 7 Gourlitz Cluster (2015) ZAF 8 Magaliesberg (2015) Zimbabwe - le Zimbabwe - Zimbabwe - Simbabwe (ZWE) ZWE 1 Middle Zambezi (2010)

On 6 to 8 December 2015, the UNESCO Office in Mexico organized the International Meeting ‘Exploring Frameworks for Tropical Forest Conservation:

managing production and consumption for sustainability’ which took place within the Framework of the United Nations 2030 Agenda for Sustainable Development at the Instituto de Ecología (INECOL) in Xalapa, Veracruz, Mexico.

After a two year process two year process of consultations and negotiations,

conducted by the United Nations, the post-2015 development framework that

will succeed the Millennium Development Goals (MDGs) was agreed upon by the 193 Member States on 2 August 2015 and adopted in September 2015 by world leaders at the Sustainable Development Summit in New York. This ambitious and forward-thinking document titled, Transforming Our World: The 2030 Agenda for

Sustainable Development, features 17 new Sustainable Development Goals (SDGs) and 169 targets that aim to end poverty, promote prosperity and people’s well-

9

TRAN RÉSER RESER GREN Poland A T Poland B K Germa C V Poland D E Roma E D Benin, F W Mauri G D Moroc H R d Portug I G El Salv J Tr Poland K W Croati L M France M Alban N O Portug O M P

ISO 31 codigo

„ext.“ dato d „ren.“ dato d

The bo used o accept Les fro sur ce l‘accep Las fro este m oficial Die au zeichn Anerk

Inform

Figure 3. UNESCO World Heritage sites in the Humid Tropics. © UNESCO Office in Mexico.

being while protecting the environment by 2030. The SDGs place great emphasis on the integration of the economic, social and environmental dimensions of sustainability – an emphasis that can and should be especially applied on a micro scale to conservation and sustainable development efforts in the tropical forest biome, which is what provided the impetus behind this meeting. Among the 17 SDGs, the following can be applied directly to the tropical forest biome and fell within the context of this meeting’s goals: Goal 6. Ensure availability and sustainable management of water and sanitation for all; Goal 12. Ensure sustainable consumption and production patterns; Goal 13. Take urgent action to combat climate change and its impacts; Goal 15. Protect, restore and promote sustainable use of terrestrial ecosystems, sustainably manage forests, combat desertification, and halt and reverse land degradation and halt biodiversity loss; Goal 16. Promote peaceful and inclusive societies for sustainable development, provide access to justice for all and build effective, accountable and inclusive institutions at all levels. The main objective of the ‘Exploring Frameworks for Tropical Forest Conservation’ meeting was to provide an international, interdisciplinary and interactive forum for the sharing and synthesis of research and progress in tropical forest conservation and sustainable development from social, economic and environmental perspectives.

10

Under the umbrella of that main objective, the meeting also pursued a few specific objectives which included the facilitation of the spread of knowledge of the state of the art of tropical forest research, conservation, and sustainable development on a global scale between Latin America and the Caribbean, Asia Pacific and Africa with the participation of national, regional and international scholars; the discussion and dissemination of environmental policies that govern today in the context of conservation and sustainable development in tropical forest regions; offering a platform to discuss the priorities, goals and methodologies of the three dimensions of sustainable development to discover where they might intersect, explore common goals and propose cross-cutting methodologies; and developing a preliminary framework that integrates the environmental, social, and economic dimensions of tropical forest conservation and sustainability which can serve as model for ‘best practice’ on an international level with a focus on applicability, adaptability and implementability. The following volume explores the history of the tropical forest from their first human encounter to the modern day anthropomorphized environments from archaeological, anthropological, ecological and biological perspectives, among others, which highlight both the importance of past populations and modern day local and indigenous communities to the conservation and sustainable use of the natural and cultural heritage that lies within. I would like to extend my sincerest gratitude to the Instituto de Ecología (INECOL) in Xalapa, the Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO) and the Gesellschaft für Internationale Zusammenarbeit (GIZ) GmbH whose collaboration led to the success of the Meeting. I would also like to thank all of the international and national experts whose commitment and willingness to share their experiences and expertise led to the fruitful dialogue which became the basis for this volume. Without these parties, this publication would not be possible.

Nuria Sanz Head and Representative UNESCO Office in Mexico

11

CHAPTER 1

Tropical Rainforests as Long-Established Anthropomorphized Landscapes

Robin Dennell University of Exeter, UK

Abstract

Resumen

The following article explores tropical forests in

El siguiente artículo explora a los bosques tropicales

various regions of the world and provides evidence

en varias regiones del mundo y proporciona eviden-

to contest popular misconceptions about tropical

cias para combatir los populares y erróneos conceptos

rainforests. The first being the narrative of a pristine,

sobre las selvas tropicales. El primero es el relato de un

‘natural landscape’; the second, that rainforests are

“paisaje natural” virgen; el segundo, aborda la unifor-

uniform across the board; and third, that Pleistocene

midad de los bosques tropicales en todos los ámbitos,

inhabitation of rainforests was highly unlikely.

y el tercero expone la poca probabilidad de la ocupa-

Evidence from Africa, China and South East Asia

ción de los bosques tropicales durante el Pleistoceno.

address these misconceptions and presents, in some

Las evidencias en África, China y el sureste de Asia

cases, a contrary narrative. In this chapter, I briefly

abordan estos conceptos erróneos y presentan, en al-

summarize this evidence and place it in its wider

gunos casos, una narrativa contraria. En este capítulo,

context.

resumiré brevemente esta evidencia y la situaré en su contexto más amplio.

14

Tropical Rainforests as LongEstablished Cultural Landscapes Introduction Tropical rainforests cover c. 10% of the Earth’s land surface, but are thought to contain half the world’s species of plants and animals. As with any other botanical

community, tropical rainforests repeatedly expanded and contracted, fragmented and recombined in response to the climatic shifts that have characterized the last 2.5 million years of the Pleistocene. Their histories are pieced together by analyses of pollen, phytoliths and lake and coastal marine sediments; other proxy indicators of rainforest histories are extant birds, lizards, mammals and genetic analyses of modern rainforest species of plants and animals. Although tropical rainforests have not been investigated in as much detail as the vegetation in temperate latitudes, it is now clear that first, present-day rainforests have complex histories, and second, they are not good analogues of their Pleistocene predecessors. These two points can be seen in a variety of regional examples of Pleistocene tropical rainforest histories. Examples include: Sémah, A. M. and Sémah, 2012; Sémah et al., 2016 for Java; Heaney, 1991 and Kershaw et al., 2007 for South-East Asia; Fei and Sun, 2004; and Zheng and Lei, 1999 for China; Bush et al., 2011; Hoorn et al., 2010; and Maslin et al., 2005 for Amazonia; Farooqui et al., 2010 for India; Hilbert et al., 2007 and Kershaw, 1994 for Australia; Cornelissen, 2013 and Dupont et al., 2000 for West and Central Africa. For those interested in the Pleistocene occupation of these regions, there are several excellent syntheses: Meltzer, 2009 for North America; Dillehay, 2000 and Moore, 2014 for South America; Barham and Mitchell, 2008 for Africa; Dennell, 2009 for Asia; and Hiscock, 2008 for Australia, as well as several contributions in this volume.

15

Tropical Rainforests as Long-Established Cultural Landscapes For pre- or non-agricultural communities of humans, living in an environment with such an amazing amount of biodiversity presented an enormous set of challenges, particularly in obtaining a regular and adequate diet. Animals that were large enough to be worth hunting tended to be largely solitary or lived in small groups; those living on the ground were often difficult to see and track, and pursuing them was impeded by dense vegetation, and those (such as monkeys) that lived in the high canopy were even more difficult to hunt. With plants, it is necessary to know which can be eaten and whether it is the roots (as with tubers such as yams), stems or fruits that are edible; some are poisonous (such as cassava) when picked and require processing by washing or boiling before they can be eaten. Rainforests can also be unhealthy places to live because of water- or insectborne diseases since open wounds can easily fester and because much of the smaller fauna can be venomous. Although rainforests have been described as the world’s largest natural pharmacy, great skill is required in knowing what plants (and which parts of them) can serve as medicines. For humans to adapt to living in rainforests was thus a major achievement, particularly for a creature that originated and long flourished in grasslands and open woodlands. There are several popular misconceptions about tropical rainforests. The first is that they are pristine, ‘natural’ landscapes, unmodified by humans. A second is that rainforests are uniform and characterized entirely by year-round rainfall, highcanopy trees, dense ground cover, unbroken vegetation and poor visibility. In fact, ‘rainforest’ is a blanket term that can hide enormous variation either altitudinally (as with montane, lowland and mangrove rainforest), spatially, depending on local soils, topography (Morley, 2000; Whitmore, 1998), and climatically, as some have dry seasons and others do not. Those that lie 5º north and south of the equator have rain year round, whereas those that are semi-tropical and between 5º and 15º north and south of the equator usually have a six-month dry season. In addition, there is always a process of clearance (through cyclones, earthquakes, natural fires, etc.) and regeneration, and these more open spaces afforded opportunities for humans in rainforests. The third misconception is the most important one in this volume, and that is, the notion that foragers could not inhabit rainforests unless they could trade or exchange with neighbouring agricultural communities. The stark implication of this viewpoint was that pre-agricultural, Pleistocene inhabitation of rainforests was highly unlikely. This view dates back to the 1980s, when (among others) Bailey et al. (1989) and Headland (Headland and Reid, 1989) independently and then jointly (Bailey and Headland, 1991) argued that there was no evidence that human foragers pre-dated agriculturalists in rainforests and made the entirely legitimate point that the fact that an archaeological site was located today in rainforest did not mean that the site lay in rainforest when it was originally occupied. As will be seen below, that challenge has now been met in some cases. In South and South-East Asia, rainforests have been extensively inhabited by humans for at

16

least 40,000 years, and perhaps even longer. In this chapter, I briefly summarize this evidence and place it in its wider context. Key sources for those interested in more detailed information are referred to Mercader (2002a, 2000b), Roberts and Petraglia (2015) and the synthesis of South-East Asia in the Pleistocene and Early Holocene by Rabett (2012).

How Early Did Humans (or Their Ancestors) Inhabit Rainforests? The earliest claims are from China and West/Central Africa; in both cases, the key issue is whether the archaeological or palaeoanthropological evidence is contemporaneous with a rainforest environment.

China One of the most fascinating animals to have inhabited South China in the Pleistocene is Gigantopithecus, which may have been 10’ (3 metres) tall judging from the size of its lower jaw (Ciochon et al. 1991, p. 166), and that probably became extinct in the late Middle Pleistocene. It was first recognized from its teeth in Chinese drugstores in the 1930s (as fossils formed an important part of traditional Chinese medicine). Later, its teeth were recovered from caves in South China and Vietnam in association with rainforest taxa such as Stegodon (a

type of extinct elephant) and Ailuripoda (panda). The teeth of Homo erectus were also found in the same layers as Gigantopithecus at some caves, and this association

gave rise to the idea that H. erectus had also inhabited rainforest. On the basis that

H. erectus and Gigantopithecus teeth were in the same layer at the cave of Tham Khuyen, Vietnam, that was dated at 475 Ka (Ciochon et al., 1996), it was thought that H. erectus had inhabited rainforest almost half a million years ago. This now

seems unlikely: the fact the teeth of H. erectus were in the same layer as those of Gigantopithecus does not necessarily mean that they were contemporaries, still

less that they lived in the same environments. Cave deposits can take thousands of years to accumulate, so Gigantopithecus and H. erectus might not have been in the same environment at the same time. Over the course of the Pleistocene,

rainforests expanded and contracted, so although Gigantopithecus was almost

certainly a forest dweller, H. erectus may have lived in the same area but only when the main vegetation was open woodland or even grassland. In a recent review, Ciochon (2009) concluded that this was likely to have been the case.

Africa Julio Mercader (2002a, 2000b) has been particularly persistent in arguing that hominins have had a long association with rainforests in West and Central Africa. For the earliest periods, the evidence is circumstantial and based on the occurrence of Acheulean and later, Lupemban and Sangoan stone artefacts in areas that are today rainforest, or may have been so in the past. Unfortunately, the sites where these artefacts are found are difficult to date because of bio-turbation by animals and insects and often have little palaeo-environmental evidence, making it difficult to demonstrate that the artefacts were made and used in rainforest. Additionally, the logistic difficulties of conducting fieldwork (for example, there is little ground visibility) in the African rainforests have been compounded by the lack of security

17

Tropical Rainforests as Long-Established Cultural Landscapes and political instability in many of these regions. A further factor is that the open landscapes of the modern African savannah in the Rift Valley and adjacent regions are perceived by many field archaeologists as being easier for fieldwork and far more likely to produce results that will lead to further successful grant applications and better career prospects. So far, the earliest, reasonably unambiguous evidence for humans in African rainforests dates from the Late Stone Age, i.e. after c. 40,000 years ago (see, for example, Cornelissen, 2013).

Java In Java, there were major changes in vegetation and fauna c. 125,000 years ago at the start of the last interglacial, when the late Middle Pleistocene Ngandong open woodland fauna was replaced by the Punung rainforest fauna that included orangutan, tapir and sun bear. Storm and colleagues (2005) claimed that a premolar from the caves Punung I and II (and in a fauna dating to the last interglacial) was that of Homo sapiens, and dated to c. 100 Ka. This claim implies that our species

inhabited the rainforests of South-East Asia at a very early date. However, the dating and identification of this tooth are in doubt (Bacon et al., 2008; Barker et al., 2007), and the latest assessment is that the premolar is at least as likely to be from late H.

erectus as from early Homo sapiens (Polanski et al., 2016). We should note, however, that there is no clear evidence from Africa and Asia that H. erectus inhabited rain

forest, and therefore “the presence of H. sapiens in a rain forest environment is more likely than occupation of this habitat by H. erectus” (Sémah and Sémah, 2012, p. 124) In sum, current archaeological and human skeletal evidence from South and

Southeast Asia clearly indicate that humans lived in rainforests after 45,000 years ago, but so far, there is no unequivocal indication that earlier types of humans were doing so in Africa or Asia. Leaving aside the morphological ambiguities of the Punung molar, the overall evidence for Homo erectus suggests that it lived in open woodlands or grasslands, and avoided rainforests.

Evidence for Humans in Rainforests After 50 Ka The earliest convincing evidence for humans inhabiting rainforests post-date 50 Ka are from south and South-East Asia. Similar evidence may be forthcoming from Laos and southern China.

Borneo: Niah Cave, and South-East Asia The vast Niah Cave on Borneo was investigated by Tom and Barbara Harrison in the 1950s, and re-investigated in much greater detail by a team directed by Graeme Barker. Thanks to this work, Niah has produced an enormous amount of environmental data, and as importantly, the Homo sapiens skull found by the

Harrisons in the deepest part of the cave’s stratigraphy has been accurately dated to c. 44 to 40 cal. Ka, making it one of the oldest indications of H. sapiens east of the Levant (see, for example, Barker, 2013; Barker et al., 2007).

Associated palaeoenvironmental data from sedimentology, pollen, phytoliths, starch grains and vertebrates indicates that the earliest inhabitants lived in a rainforest environment (Hunt et al., 2012). Faunal evidence shows that monkeys that inhabit

18

the high canopy were regularly killed, and starch grain from toxic plants indicates that these were collected and processed for consumption (Barton, 2005; Piper and Rabett, 2014). Although Niah has currently the earliest evidence for humans living in rainforest, the reality may have been more subtle. The local vegetation changed over time, and included grassland, as well as montane, mangrove and lowland forest. The faunal data included taxa from a mosaic environment as well as rainforest, and pollen and charcoal data indicate that the inhabitants regularly burnt vegetation to promote open habitats. This pattern, of living in mosaic environments that included rainforest as well as open environments, is one that is likely common across South-East Asia after 40 Ka. Examples are the caves of Lang Rongrien, Thailand, between 32 Ka and 44 Ka (Anderson, 1997; Mudar and Anderson, 2007), and Lang Kamnam, Thailand after 27 Ka (Shoocongdej, 2000); Punung, Java after 45 Ka (Sémah and Sémah, 2012), and caves on the Tabon Peninsula, Palawan in the southern Philippines (Pawlik et al., 2014). Rainforests may have been used prior to 45 Ka in South-East Asia. The cave of Tam Pa Ling, Laos, is potentially a good candidate for a site where rainforests were exploited before 45 Ka. It has produced human remains dated to between 46 and 63 Ka (Demeter et al., 2012 but see Pierret et al., 2012 who argue that the maximum age of Tam Pa Ling human TPL1 is only 46 Ka), but palaeoenvironmental data are currently lacking. Another potential candidate is the cave of Xiaodong, Yunnan Province, south-west China, with occupation dated at 43.5 Ka (Ji et al., 2015). Pollen data indicate that the site was in rainforest, and the faunal remains are consistent with that interpretation (Ji et al., 2015). Isotopic analysis is needed to see if these early Hoabinhians were living in rainforest.

Sri Lanka The most convincing evidence for humans living wholly in rainforests before 30 Ka is currently from Sri Lanka. Excavations of the caves of Batadomba-lena, Fa Hien-lena and Kitulgala Beli-lena have produced human remains associated with a microlithic assemblage (likely for use in hafted tools) (Lewis et al., 2015; Roberts et al., 2015) and an extensive bone industry that included points that were probably used as projectiles or in snares as early as 36 to 38 Ka (Perera et al., 2011). Faunal data indicates that semi-arboreal and arboreal primates comprise c. 70 to 80% of the mammalian assemblages. Other resources exploited at Batadomba-lena include mouse deer, giant squirrel, mongoose, jungle cat and civet, as well as Canarium sp. nuts and starchy rainforest plants (Perera et al., 2011), all of which imply dedicated

rainforest subsistence. In a ground-breaking investigation, Roberts and colleagues (2015) showed from stable carbon and oxygen isotopic analysis of human and other animal bones from these caves that human diet was overwhelmingly from rainforest foods. Moreover, this lifestyle was maintained from 36 Ka through the rest of the Pleistocene and into the Holocene up to when agriculture was introduced c. 3,000

years ago. Sri Lanka thus has the earliest evidence for life in the rainforest, as well as the longest maintained tradition of doing so.

19

Tropical Rainforests as Long-Established Cultural Landscapes

New Guinea, Melanesia Another area where rainforests were used at an early date is in the Ivane Valley in highland New Guinea, where archaeological sites dated to between 43 and 49 Ka have been found at 2,000 m above sea level. Waisted stone axes dated to c. 40 Ka are thought to have been used for clearing forest, and starch grains indicate the use of forest plants such as Dioscorea yam species and Pandanus nuts from as early as 40

to 30 Ka (Summerhayes and Ford, 2014; Summerhayes et al., 2010). In west New Britain in the Bismarck Islands, tropical forest resources were being used as early as 35.5 Ka by colonists who had arrived there by boat (Pavlides and Gosden, 1994).

The Wider Context The evidence for humans adapting to rainforests by 40 Ka and perhaps earlier in east and South-East Asia is consistent with other evidence for far-reaching, novel adaptations around this time. Recent evidence, particularly from east and SouthEast Asia, is demonstrating remarkable innovations by humans after 40 Ka and perhaps even earlier. These include the regular use of sea craft by which groups were reaching places that had previously been inaccessible. For example, in the Philippines at 67 Ka, as seen by human/hominin remains at Callao Cave (although the identity of the inhabitant is unclear) (Mijares et al., 2010); Sahel (the conjoined landmass of Australia, Tasmania and New Guinea) by 50 Ka and perhaps by 60 Ka (Roberts et al., 1994); Timor, for example, was colonised by sea by 42 Ka (O’Connor, 2007); Japan (PalaeoHonshu) by 38 Ka (Kudo and Kumon, 2012); and Okinawa by 32 Ka (Kaifu et al., 2015). One recent spectacular discovery is that the painted rock art on Sulawesi, dated at 36 to 40 Ka, is as ancient as the earliest cave art in Western Europe (Aubert et al., 2015). Japanese obsidian was also being obtained from off-shore islands and then exchanged over networks that spanned at least 1,000 km (Kuzmin et al., 2002; Kuzmin, 2006).

Implications for the Americas The first arrivals to reach the Americas probably entered via the exposed and enormous continental shelf that made up Beringia, which linked north-east Asia and Alaska across what is now the Bering Strait. These immigrants would not have had any experience of rainforest environments, so in a real sense, they did not arrive with direct knowledge of how to survive in one. What they did arrive with was the problem-solving abilities and adaptability of modern humans; thus, when they encountered rainforests in Central and South America, they had the mental capacity to learn how to use them. In Amazonia, their main adaptation lay in utilizing an astonishing variety of plant resources, and creating agricultural landscapes. In Amazonia and Central America, several millennia of foraging after ca. 11,500 BP (Dillehay 2009) preceded the emergence of agricultural landscapes characterized by extensive field and drainage systems (marked by the terra preta or Amazonian Dark Earths) over the last 2000-3000 years. The earliest of these field systems date to ca. 450 BC, but some may be earlier. It has been estimated that these may cover up to 0.3% of the Amazon Basin, or ca. 18,000 sq.km., with some covering 500 hectares (Moore 2014, p. 155). A striking aspect of the archaeological record from the Amazonian rainforest is that there was not a simple pathway from

20

foraging to farming: a wide variety of local and introduced plants were cultivated for millennia before they became morphologically domesticated, and without the creation and maintenance of field and drainage systems.

Conclusions The most important point to have emerged from archaeological research in rainforests is that these are not “not wild, pristine expanses of raw nature, but rather are – to varying degrees – anthropogenic landscapes” (Moore 2014, p. 153) that have been modified by humans for millennia, whether in Amazonia, Central America, South and Southeast Asia, or Africa. At present, the earliest clear evidence of life in the rainforest is from South and South-East Asia, particularly from Niah Cave, Borneo, and the caves of Sri Lanka. Evidence here strongly implies that Homo sapiens, and not its predecessors, were the

first to inhabit rainforest. Earlier claims for living in rainforest in Middle Stone Age Africa and Asia before 100,000 years ago are currently unsubstantiated because of the difficulties of dating artefactual material and demonstrating that it was used in a rainforest. In Australia and the Americas, the ability to live in rainforests was rediscovered by the first colonisers, and life in rainforests in these continents developed along its own idiosyncratic pathway, in areas with a completely different flora and fauna. The adaptation to living in rainforests is as significant as that of living in the Arctic and provides an outstanding example of human adaptability after 40 to 50 Ka. Although the earliest sites in rainforests – whether in south or South-East Asia, Africa or the Americas – are not spectacular, their Outstanding Universal Value lies in the demonstration of human adaptability in the most challenging of circumstances.

References Anderson, D. D. 1997. Cave archaeology in Southeast Asia. Geoarchaeology, Vol. 12, No. 6, pp. 607–38.

Aubert, M., Brumm, A., Ramli, M., Sutikna, T., Saptomo, E. W., Hakim, B., Morwood, M., van den Bergh, G. D., Kinsley, L. and Dosseto, A. 2014. Pleistocene cave art from Sulawesi, Indonesia. Nature, Vol. 514, pp. 223–27. Bacon, A. M., Demeter, F., Duringer, P., Helm, C., Bano, M., The Long, V. and Thuy, N. et al. 2008. The Late Pleistocene Duoi U’Oi Cave in Northern Vietnam: Palaeontology, sedimentology, taphonomy and palaeoenvironments. Quaternary Science Reviews, Vol. 27, pp. 1627–54.

Bailey, R. C., and Headland, T. N. 1991. The tropical rain forest: is it a productive environment for human foragers? Human Ecology, Vol. 19, No. 2, pp. 261–85. Bailey, R., Head, G., Jenike, M., Owen, B., Rechtman, R. and Zechenter, E. 1989. Hunting and gathering in tropical rain forest: is it possible? American Anthropologist, Vol. 91, pp. 59–82.

21

Tropical Rainforests as Long-Established Cultural Landscapes Barham, L. and Mitchell, P. 2008. The First Africans: African Archaeology from the Earliest Toolmakers to Most Recent Foragers. Cambridge, UK, Cambridge University Press.

Barker, G. (ed.). 2013. Rainforest foraging and farming in Island Southeast Asia, Vol. 1: The Archaeology of the Niah Caves, Sarawak. Cambridge, UK, McDonald Institute for Archaeological Research.

Barker, G., Barton, H., Bird, M., Daly, P., Datan, I., Dykes, A. and Farr, L. et al. 2007. The ‘Human Revolution’ in lowland tropical Southeast Asia: The antiquity and behavior of Anatomically Modern Humans at Niah Cave (Sarawak, Borneo). Journal of Human Evolution, Vol. 52, pp. 243–61. Barton, H. 2005. The case for rainforest foragers: The starch record at Niah Cave, Sarawak. Asian Perspectives, Vol. 44, No. 1, pp. 56–72. Bush, M. B., Gosling, W. D. and Colinvaux, P. A. 2011. Climate and vegetation change in the lowlands of the Amazon Basin. M. Bush, J. Flenley and W. Gosling (eds), Tropical Rainforest Responses to Climatic Change. Berlin, Springer Praxis Books, pp. 61-84.

Ciochon, R. L. 2009. The mystery ape of Pleistocene Asia. Nature, Vol. 459, pp. 910–911.

Ciochon, R., The Long, V., Larick, R., González, L., Grün, R., de Vos, J., Yonge, C., Taylor, L., Yoshida, H. and Reagan, M. 1996. Dated co-occurrence of Homo erectus and Gigantopithecus from Tham Khuyen Cave, Vietnam. Proceedings of the National Academy of Science US, Vol. 93, No. 7, pp. 3016–20.

Ciochon, R. L., Olsen, J. and James, J. 1991. Other Origins: The Search for the Giant Apei n Human Prehistory. London, Victor Gollancz Ltd.

Ciochon, R., The Long, V., Larick, R., González, L., Grün, R., de Vos, J., Yonge, C., Taylor, L., Yoshida, H. and Reagan, M. 1996. Dated co-occurrence of Homo erectus and Gigantopithecus from Tham Khuyen Cave, Vietnam. Proceedings of the National Academy of Science US, Vol. 93, No. 7, pp. 3016–3020.

Cornelissen, E. 2013. Hunting and gathering in Africa’s tropical forests at the end of the Pleistocene and in the Early Holocene. P. Mitchell and P. Lane (eds), The Oxford Handbook of African Archaeology. Oxford, UK, Oxford University Press, pp. 403-417.

Demeter, F., Shackelford, L. L., Bacon, A. M., Duringer, P., Westaway, K., Sayavongkhamdy, T. and Braga, J. et al. 2012. Anatomically modern human in Southeast Asia (Laos) by 46ka. Proceedings of the National Academy of Sciences of the United States of America, Vol. 109, pp. 14375–14380.

22

Dennell, R. W. 2009. The Palaeolithic Settlement of Asia. Cambridge, UK, Cambridge University Press.

Dillehay, T. 2000. The Settlement of the Americas. New York, Basic Books. Dupont, L. M., Jahns, S., Marret, F. and Shi Ning. 2000. Vegetation change in equatorial West Africa: time-slices for the last 150 Ka. Palaeogeography, Palaeoclimatology, Palaeoecology, Vol. 155, pp. 95–122.

Farooqui, A., Ray, J. G., Farooqui, S. A., Tiwari, R. K. and Khan, Z. A. 2010. Tropical rainforest vegetation, climate and sea level during the Pleistocene in Kerala, India. Quaternary International, Vol. 213, pp. 2–11. Fei, H. and Sun, X. 2004. Palynological record during the Pleistocene (between 1.05 Ma and 0.36 Ma) from ODP1144, northern South China Sea. Science in China Series D: Earth Sciences, Vol. 47, pp. 30-36.

Headland, T. N. and Reid, L. A.1989. Hunter-gatherers and their neighbors from prehistory to the present. Current Anthropology, Vol. 30, pp. 43–66. Heaney, L. R. 1991. A synopsis of climatic and vegetational change in Southeast Asia. Climate Change, Vol. 9, pp. 53-61. Hilbert, D. W., Graham, A., Hopkins, M. S. 2007. Glacial and interglacial refugia within a long-term rainforest refugium: The Wet Tropics Bioregion of NE Queensland, Australia. Palaeogeography, Palaeoclimatology, Palaeoecology, Vol. 251, pp. 104–18.

Hiscock, P. 2008. Archaeology of Ancient Australia. Abingdon, UK, Routledge. Hoorn, C., Wesselingh, F. P., ter Steege, H., Bermudez, M. A. and Mora, A. et al. 2010. Amazonia through time: Andean uplift, climate change, landscape evolution, and biodiversity. Science, Vol. 330, pp. 927-31. Hunt, C. O., Gilbertson, D. D. and Rushworth, G. 2012. A 50,000-year record of Late Pleistocene tropical vegetation and human impact in lowland Borneo. Quaternary Science Reviews, Vol. 37, pp. 61–80. Ji, W., Kuman, K., Clarke, R.J., Forestier, H., Li, Y., Ma, J., Qiu, K., Li, H. and Wu, Y. 2016. The oldest Hoabinhian technocomplex in Asia (43.5 ka) at Xiaodong rockshelter, Yunnan Province, southwest China. Quaternary International, Vol. 400, pp. 166-174.

Kaifu, Y., Fujita, M., Yoneda, M. and Yamasaki, S. 2015. Pleistocene seafaring and colonization of the Ryuku Islands, Japan. Y. Kaifu, T. Goebel, H. Sato, and

23

Tropical Rainforests as Long-Established Cultural Landscapes A. Ono, (eds), Emergence and diversity of Modern Human Behavior in Paleolithic Asia. Texas, Texas A&M University Press, College Station, pp. 345-361.

Kershaw, A. P. 1994. Pleistocene vegetation of the humid tropics of northeastern Queensland, Australia. Palaeogeography, Palaeoclimatology, Palaeoecology, Vol. 109, pp. 399-412.

Kershaw, A. P., Kaars, S. van der, Flenley, J. R. 2007. The Quaternary history of far eastern rainforests. Tropical Rainforest Responses to Climatic Change. Springer Praxis Books, pp. 77-115.

Kudo, Y. and Kumon, F. 2012. Paleolithic cultures of MIS 3 to MIS 1 in relation to climate changes in the central Japanese islands. Quaternary International, Vol. 248, pp. 22-31.

Kuzmin, Y. V. 2006. Recent studies of obsidian exchange networks in prehistoric Northeast Asia. D. E. Dumond and R. L. Bland (eds), Archaeology in Northeast Asia: On the Pathway to Bering Strait. Oregon, Museum of Natural History, pp. 61-72.

Kuzmin, Y. V., Glascock, M. D. and Sato, H. 2002. Sources of archaeological obsidian on Sakhalin Island (Russian Far East). Journal of Archaeological Science, Vol. 29, pp. 741–49.

Lewis, L., Perera, N. and Petraglia, M. 2014. First technological comparison of Southern African Howiesons Poort and South Asian microlithic industries: An exploration of inter-regional variability in microlithic assemblages. Quaternary International, Vol. 350, pp. 7–25.

Maslin, M., Malhi, Y., Phillips, O. and Cowling, S. 2005. New views on an old forest: assessing the longevity, resilience and future of the Amazon rainforest. Transactions of the Institute of British Geographers, Vol. 30, pp. 477–99. Meltzer, D. J. 2009. First People in a New World. Berkeley, Calif., University of California Press.

Mercader, J. 2002a. Forest People: The role of African rainforests in human evolution and dispersal. Evolutionary Anthropology, Vol. 1, pp. 117–24. Mercader, J., ed. 2002b. Under the Canopy: The Archaeology of Tropical Rain Forests. Piscataway, NJ, Rutgers University Press.

Mijares, A. S. B., Détroit, F., Piper, P., Grün, R., Bellwood, P., Aubert, M., Champion, G., Cuevas, N., De Leon, A. and Dizon, E. 2010. New evidence for a 67,000-year-old human presence at Callao Cave, Luzon, Philippines. Journal of Human Evolution, Vol. 59, pp. 123–32.

24

Moore, J. D. 2014. A Prehistory of South America: Ancient Cultural Diversity on the Least Known Continent. Boulder, Colo., University Press of Colorado.

Morley, R. J. 2000. Origin and Evolution of Tropical Rain Forests. New York, Wiley & Sons.

Mudar, K. M. and Anderson, D. 2007. New evidence for Southeast Asian Pleistocene foraging economies: Faunal remains from the early levels of Lang Rongrien rockshelter, Krabi, Thailand. Asian Perspectives, Vol. 46, pp. 298–334. doi:10.1353/asi.2007.0013.

O’Connor, S., 2007. New evidence from East Timor contributes to our understanding of earliest colonisation east of the Sunda Shelf. Antiquity, Vol. 81, pp. 523-35. Pavlides, C. and Gosden. C. 1994. 35,000-year-old sites in the rainforests of West New Britain, Papua New Guinea. Antiquity, Vol. 68, pp. 604–10. Pawlik, A. F., Piper, P. J. and Mijares, A. S. B. 2014. Modern humans in the Philippines: Colonization, subsistence and new insights into behavioural complexity. R. Dennell and M. Poor (eds), Southern Asia, Australia and the Search for Human Origins. Cambridge, UK, Cambridge University Press, pp. 135–47.

Perara, N., Kourampas, N., Simpson, I. A., Deraniyagala, S. U., Bulbeck, D., Kamminga, J., Perera, J., Fuller, D. Q., Szabo, K. and Oliviera, N. V. 2011 People of the ancient rainforest: Late Pleistocene foragers at the Batadomba-lena rockshelter, Sri Lanka. Journal of Human Evolution, Vol. 61, pp. 254-69. Pierret, A., Zeitoun, V. and Forestier, H. 2012. Irreconcilable differences between stratigraphy and direct dating cast doubts upon the status of Tam Pa Ling fossil. Proceedings of the National Academy of Sciences of the United States of America, Vol. 109, No. 51, E3523.

Piper, P. J. and Rabett, R. J. 2014. Late Pleistocene subsistence strategies in Island Southeast Asia and their implications for understanding the development of modern human behaviour. R. Dennell and M. Poor (eds), Southern Asia, Australia and the Search for Human Origins. Cambridge, UK, Cambridge University Press, pp. 118–34.

Polanski, J. M., Marsh, H. E. and Maddux, S. D. 2016. Dental size reduction in Indonesian Homo erectus: Implications for the PU-198 premolar and the appearance of Homo sapiens on Java. Journal of Human Evolution, Vol. 90, pp. 49-54.

Rabett, R. 2012. Human Adaptation in the Asian Palaeolithic: Hominin Dispersal and

Behaviour during the Late Quaternary. Cambridge, UK, Cambridge University Press.

25

Tropical Rainforests as Long-Established Cultural Landscapes Roberts, P. and Petraglia, M. D. 2015. Pleistocene rainforests: barriers or attractive environments for early human foragers? World Archaeology, Vol. 47, No. 5. Roberts, P., Boivin, N. and Petraglia, M. D. 2015a. The Sri Lankan ‘microlithic’ tradition C. 38,000 to 3000 years ago: Tropical technologies and adaptations of Homo sapiens at the southern edge of Asia. Journal of World Prehistory, Vol. 28, No. 2, pp. 69-112.

Roberts, P., Perera, N., Wedage, O., Deraniyagala, S., Perera, J., Eregama, S., Gledhill, A., Petraglia, M. D. and Lee-Thorpe, J. 2015b. Direct evidence for human reliance on rainforest resources in late Pleistocene Sri Lanka. Science, Vol. 347, pp. 1246-49.

Roberts, R. G., Jones, R., Spooner, N. A., Head, M. J., Murray, A. S., Smith, M.A., 1994. The human colonisation of Australia: optical dates of 53,000 and 60,000 years bracket human arrival at Deaf Adder Gorge, Northern Territory. Quaternary Geochronology (Quaternary Science Reviews), Vol. 13, pp. 575-83.

Sémah, A.-M. and Sémah, F. 2012. The rain forest in Java through the Quaternary and its relationships with humans (adaptation, exploitation and impact on the forest). Quaternary International, Vol. 249, pp. 120-128. Sémah, A.-M., Sémah, F., Moigne, A.-M., Ingicco, T., Purnomo, A., Simanjuntak, T. and Widianto, H. 2016. The palaeoenvironmental context of the Palaeolithic of Java: A brief review. Quaternary International, Vol. 416, pp. 38-45. Shoocongdej, R. 2000. Forager mobility organization in seasonal tropical environments of Western Thailand. World Archaeology, Vol. 32, pp.14-40. Storm, P., Aziz, F., de Vos, J., Kosasih, D., Baskoro, S., Ngaliman, and van den Hoek Ostende, L. W. 2005 Late Pleistocene Homo sapiens in a tropical rainforest fauna in East Java. Journal of Human Evolution, Vol. 49, pp. 536–45.

Summerhayes, G. R. and Ford, A. 2014. Late Pleistocene colonisation and adaptation in New Guinea: Implications for modeling modern human behaviour. R. Dennell and M. Poor (eds), Southern Asia, Australia and the Search for Human Origins. Cambridge, UK, Cambridge University Press, pp. 213–27.

Summerhayes, G. R., Leavesley, M., Fairbairn, A., Mandui, H. Field, J. Ford, A. and Fullagar, R. 2010. Human adaptation and plant use in Highland New Guinea 49,000 to 44,000 years ago. Science, Vol. 330, pp. 78–81. Whitmore, T. C. 1998. An Introduction to Tropical Rainforests, 2nd edn. Oxford, UK, Oxford University Press.

26

Zheng, Z. and Lei, Z.-Q. 1999. A 400,000 year record of vegetational and climatic changes from a volcanic basin, Leizhou Peninsula, southern China. Palaeogeography, Palaeoclimatology, Palaeoecology, Vol. 145, pp. 339–62.

27

Patrick Roberts School of Archaeology, Research Laboratory for Archaeology and the History of Art, University of Oxford, UK

Abstract

Resumen

Until relatively recently, archaeologists argued that

Hasta hace relativamente poco tiempo, los arqueólo-

intensive rainforest occupation and exploitation by

gos solían argumentar que la intensa ocupación y ex-

our own species began at the start of the Holocene (c.

plotación de los bosques lluviosos por parte de nuestra

11 Ka). This was supported by anthropological and

especie empezó al inicio del periodo Holoceno (hace

ecological arguments that suggested tropical rainforests

alrededor de 11 mil años). Esta información fue respal-

presented difficulties of navigation and limited protein

dada por argumentos antropológicos y ecológicos que

and carbohydrate resources, requiring sophisticated

sugerían que los bosques lluviosos tropicales imponían

human technologies and extraction strategies to be

dificultades para la navegación y que limitaban los re-

effectively exploited. Nevertheless, over the last decade,

cursos de proteínas y de carbohidratos, requiriendo,

archaeological research in Africa, Southeast Asia, and

así, una explotación efectiva de tecnologías humanas

Melanesia has demonstrated that humans were able

sofisticadas y estrategias de extracción. Con todo, en

to inhabit and exploit rainforest ecologies at least as

la última década, investigaciones arqueológicas efec-

early as 45 Ka (thousand years ago). Here, I review

tuadas en África, en el Sudeste Asiático y en Melanesia

South Asia as a particularly interesting case study in

han demostrado que el hombre fue capaz de habitar

the growing literature of Late Pleistocene human

y explotar la ecología de los bosques lluviosos segu-

rainforest adaptations. The earliest Homo sapiens

ramente desde hace 45 mil años atrás. Aquí, analizo

fossils in this region occur in the modern-day Wet

Asia del Sur como un caso de estudio particularmente

Zone rainforests of Sri Lanka c. 36 Ka, accompanied

interesante en la cada vez mayor literatura acerca de

by early microlith toolkits, bone tool assemblages

las adaptaciones del hombre del Pleistoceno tardío a

and complex hunting strategies that persist from this

los bosques lluviosos. Los primeros fósiles del Homo

early period of rainforest colonisation down to the Sri Lankan Iron Age c. 3 Ka. In reviewing early human

sapiens de esta región se encontraron (hace aproximadamente 36 mil años) en las junglas y bosques lluviosos

rainforest interactions in Sri Lanka, I also highlight the

de lo que hoy es Sri Lanka, junto con las primeras cajas

potential of stable carbon and oxygen isotope analysis

de herramientas líticas, ensamblajes con utensilios he-

of fossil human and faunal tooth enamel to provide

chos de huesos y estrategias de cacería complejas que

further detail in studies of prehistoric human rainforest

prevalecieron desde aquel periodo temprano de colo-

adaptations.

nización de la jungla hasta la Edad del Hierro en Sri Lanka (hace aproximadamente 3 mil años). Al revisar las primeras interacciones humanas con el bosque lluvioso en Sri Lanka, para brindar mayores detalles al estudio de la adaptación del hombre prehistórico a la selva tropical, resalto el potencial del análisis de isótopos estables de carbono y de oxígeno de fósiles humanos, así como del esmalte de dientes de animales.

28

Early Human Adaptation to Late PleistoceneHolocene Rainforests in South Asia Introduction Until relatively recently, archaeologists argued that intensive rainforest occupation and exploitation by our own species began at the start of the Holocene (c. 11 Ka) (for example, Gamble, 1993). This was supported by anthropological and ecological arguments that suggested tropical rainforests presented difficulties of navigation, and limited protein and carbohydrate resources, requiring sophisticated human technologies and extraction strategies to be effectively exploited (Bailey and Headland, 1991). Nevertheless, over the last decade, archaeological research in Africa, South-East Asia and Melanesia has demonstrated that humans were able to inhabit and exploit rainforest ecologies at least as early as 45 Ka (thousand years ago) (Mercader, 2002a,b; Barker et al., 2007; Summerhayes et al., 2010). Here, I review South Asia as a case study of Outstanding Universal Value to World Heritage in the growing literature of Late Pleistocene human-rainforest adaptations. South Asia, and more particularly Sri Lanka, has yielded some of the earliest dated fossil evidence for Homo sapiens (c. 36 Ka) in a modern rainforest context beyond Africa.

This fossil evidence is accompanied by early and stable microlithic toolkits, bone technologies, personal ornamentation and complex rainforest hunting strategies that persist until the appearance of the Iron Age in the tropical forests of the region (c. 3 Ka). In reviewing early human rainforest interactions in Sri Lanka, I also highlight the potential of stable carbon and oxygen isotope analysis of fossil human and faunal tooth

29

Early Human Adaptation to Late Pleistocene-Holocene Rainforests in South Asia enamel to provide further detail in studies of prehistoric human rainforest adaptations. The majority of existing research into early human rainforest adaptations around the globe has relied on indirect pollen, archaeobotanical and archaeozoological palaeoenvironmental records, of unknown resolution and catchment, in association with archaeological sequences. By contrast, stable isotope methodologies have the potential to provide direct insight into human rainforest resource reliance. Researchers from the University of Oxford and the Department of Archaeology, Government of Sri Lanka collaborated to apply stable carbon and oxygen isotope analysis to Late Pleistocene-Holocene human fossils in Sri Lanka. I propose that this methodology should be applied to fossil human and faunal remains uncovered in other regions of the world where human-rainforest relationships of global importance are identified. The resulting compilation of isotopic datasets across time and space promises to enrich international understanding of the long-term interaction of our species with the most ecologically diverse environments on our planet.

The Sri Lankan Microlithic Tradition: Late Pleistocene Rainforest Specialization? In Sri Lanka, the ‘Microlithic’ characterizes a tradition of small stone tools, less than 2 cm long and sometimes fashioned into geometric microliths (Deraniyagala, 1992; Lewis et al., 2014; Roberts et al., 2015a) (Figure 1). The earliest Microlithic tradition comes from the three Wet Zone rainforest cave and rockshelter sites of Fa Hien-lena, Batadomba-lena and Kitulgala Beli-lena at c. 38 Ka, 36 Ka and Figure 1. Examples of quartz microliths excavated from Batadomba-lena Layer 7c (36 Ka) recently published in Lewis et al. (2014). © Laura Lewis.

30

30 Ka, respectively, and is associated with the earliest fossil evidence of Homo

sapiens in Sri Lanka at Fa Hien-Lena (c. 38-36 Ka) and Batadomba-lena (c. 36 Ka)

(Deraniyagala, 1992; Perera, 2010, Perera et al., 2011; Roberts et al., 2015a) (Figure 2). Other material culture traits across these sites include an early and stable osseous

technology of ‘double’ and ‘single’ points (Deraniyagala, 1992; Perera, 2010), the use of red and yellow ochre pigment (Perera et al., 2011) and evidence for longdistance exchange (Perera, 2010). These material features and the accompanying fossil record continue in the Wet Zone of Sri Lanka into the Terminal Pleistocene and Early Holocene. From this point, the Microlithic tradition expands into other climatic zones and ecologies on the island, including coastal zones, where occasional microlithic finds are associated with shell middens along the southern and north-western portions of the island (Deraniyagala, 1992; Roberts et al., 2015a; Somadeva and Ranasinghe, 2006). The Late Pleistocene-Early Holocene Microlithic tradition record is therefore solely confined to the Wet Zone rainforests of modern-day Sri Lanka. That these inhabited ecologies were also rainforest in the past is supported by faunal records from Fa Hien-lena and Batadomba-lena that demonstrate the capture of forestdwelling Moschiola memina (mouse deer), Hystrix indica (porcupine), Parradoxurus sp. (palm cat), Viverricula indica (palm civet) and bat species (Perera et al., 2011;

Roberts et al., 2015b). The mammalian faunal records from these sites suggest that Figure 2. Wet Zone rainforest site Batadomba-lena that provides some of the earliest evidence for Homo sapiens fossils, microlith technologies and bone technologies in South Asia. © Patrick Roberts.

forest-dwelling, semi-arboreal and arboreal primates consistently make-up between 70% and 80% of the mammalian assemblage over the period of occupation (Perera, 2010; Perera et al., 2011; Wijeyapala, 1997). This proportion is unheard of in both the archaeological and anthropological literature, even with the modern-day availability of rifle technologies in many parts of the Amazon rainforest (Altherr,

31

Early Human Adaptation to Late Pleistocene-Holocene Rainforests in South Asia 2007; Lu, 2010; Papworth et al., 2013), with perhaps one exception (Fortier, 2014), and implies a well-tuned, specialized exploitation strategy. Additional evidence for the exploitation of rainforest resources comes from the abundance of Canarium sp.

nuts and the presence of abundant freshwater and forest molluscs throughout the sequence of Batadomba-lena (Perera et al., 2011). Although the early archaeology of the Microlithic tradition has led to Batadombalena, Fa Hien-lena and Kitulgala Beli-lena now being listed as a combined World Heritage site, Sinharaja Forest Reserve – inscribed in 1988 under criterias (ix) and (x) – the role these sites play in our understanding of early human-rainforest adaptations has thus far been downplayed. This is partly a result of the fact that, as in most regions of the world where Late Pleistocene human-rainforest interactions have been suggested, questions remain regarding the extent and nature of human rainforest adaptations in Sri Lanka. ‘Off-site palaeoenvironmental and palaeoclimatic data across Sri Lanka is currently sparse and fragmented, making it difficult to effectively reconstruct the composition, nature and density of tropical rainforest in the vicinity of Sri Lankan Microlithic sites through time (Perera et al., 2011; Roberts et al., 2015a). Furthermore, ‘on-site’ palaeoenvironmental evidence from molluscan and faunal remains and reconstructed taxa habitat preferences can be relatively coarse. Given the environmental and climatic diversity in Sri Lanka (Figure 3), it is possible that plants and animals could have been transported to the site from more distant ecologies. As a result, whether the Late Pleistocene Microlithic tradition represents rainforest specialization, as opposed to seasonal rainforest resource use, has remained an open question.

Stable Carbon and Oxygen Isotope Analysis of Fossil Tooth Enamel: a Test of Human Rainforest Resource Reliance Stable isotopes of an element differ slightly in their nuclear mass as a result of differences in the number of neutrons. The absolute stable isotope abundance of a sample is difficult to determine (Hayes, 1983). Consequently, isotope ratios are expressed as the relative abundance of the heavy (minor) to light (major) isotope using the delta notation in per mil. 1) δaX (‰) = (Rsample/Rstandard-1)*1000, where R is the ratio of the heavy (a) to light (b) isotope of X. The mass difference in isotopes of the same element result in small but significant variation in thermodynamic and kinetic properties (Sharp, 2006). Kinetic effects in fast, incomplete or unidirectional reactions involve lighter isotopes being more ‘mobile’ and reacting faster than heavier ones. In enzyme-driven biochemical reactions, the enzyme discriminates against the heavier isotope in the system. In equilibrium reactions, isotope effects generally relate to the effects of atomic mass on bond energy. Heavier isotopes form stronger bonds and will therefore be concentrated in phases with higher bond energy (Hoefs, 1997; Sharp, 2006). These

32

Figure 3. The location of 1. Batadomba-lena, 2. Fa Hien-lena, 3. Balangoda Kuragala and 4. Bellan-bandi Palassa relative to a) elevation b) precipitation modelled using Worldclim 1.3 records of international precipitation and shown in three arithmetic groups discussed here as the Wet Zone, Intermediate Zone and Dry Zone and c) map of the vegetation redrawn using data from Dittus (1977) and Erdelen (1988). © Patrick Roberts. fundamental fractionation phenomena cause subtle, but measurable, differences in isotope abundance in biogenic materials that can be related to biological processes such as photosynthesis and food metabolism, and environmental factors such as temperature (Koch, 1998; Sharp, 2006). Stable carbon and oxygen isotope analysis of human and faunal tooth enamel has the potential to provide a direct record of human rainforest resource reliance. Based on the δ13C distinction between the C3 and C4 photosynthetic pathways at the base of tropical foodwebs, stable carbon isotope analysis of hominin and associated

faunal tissues in the tropics and subtropics has frequently been used to assess the proportion of C3 or C4 biomass in their diets (Lee-Thorp, 2008; Levin et al., 2008;

Sponheimer et al., 2013). Within tropical forest ecologies, this distinction can be

33

Early Human Adaptation to Late Pleistocene-Holocene Rainforests in South Asia used to assess the degree of faunal reliance on versus

13

13

C-depleted C3 forest resources

C-enriched C4 plant resources available in open habitats (for example,

White et al., 2009). This division is further enhanced by the ‘canopy effect’, where

vegetation growing under a closed forest canopy is strongly depleted in 13C, due to low light and an abundance of respired CO2 (van der Merwe and Medina, 1991).

As a consequence, the tissues of forest-dwelling animals have more negative δ13C values than those spending some, or all, of their time consuming open-habitat foodstuffs (Cerling et al., 2004). For example, modern enamel from fauna living in the closed-canopy C3 Ituri Forest of the Congo-Kinshasa have a δ13C range of -26.0 to -14.1‰, relative to a range of -12.0 to 0.2‰ for herbivores in open C3/

C4 grasslands elsewhere in East Africa (Levin et al., 2008). Stable oxygen isotope

measurements from faunal enamel provide additional palaeoecological information about water and food (Koch, 2007) that may, in turn, be influenced by the degree of forest closure. In contrast to many human and faunal tissues, tooth enamel consists primarily of an inorganic fraction that is extremely resistant to post-mortem diagenesis, even in the humid environments of the tropics (Krigbaum, 2003, 2005). This has led to stable carbon and oxygen isotope analysis of tooth enamel being used to examine the forest reliance and adaptations of fauna, hominins and extinct apes back to 7 Ma (Bocherens et al., 2015; Nelson, 2003, 2007; White et al., 2009). In the context of prehistoric Homo sapiens, Krigbaum (2003, 2005) has used stable carbon and oxygen isotopes from Early to Middle Holocene human tooth enamel to elucidate

dietary changes between Neolithic human groups entering the tropical forests of South-East Asia and those foraging within forested environments during the pre-Neolithic phase. Despite this work, there has thus far been little attempt to develop a direct stable isotope record of Late Pleistocene human subsistence in any of the regions where early human rainforest occupation and exploitation has been inferred. Collaboration between the University of Oxford and the Department of Archaeology, Government of Sri Lanka sought to remedy this by applying stable carbon and oxygen isotope analysis to human and faunal tooth enamel from four Late Pleistocene-Holocene sequences in Sri Lanka. Tooth enamel samples from a variety of fauna and humans from the Terminal Pleistocene/Holocene deposits (c. 12,000 to 3,000 cal. years BP) of Fa Hienlena, Balangoda Kuragala and Bellan-bandi Palassa provided a suite of samples encompassing the full climatic and ecological spectrum of the island’s modern climate and vegetation zones (Roberts et al., 2015b) (Figure 3). Human specimens dated to c. 20,000 to 17,00 cal. years BP from Batadomba-lena, in association with a faunal sequence spanning 20,000 to 12,000 cal. years BP, provided an

earlier record of Wet Zone rainforest hunter-gatherer adaptations. δ13C and δ18O measurements from fauna from Fa Hien-lena, Balangoda Kuragala and Bellan-bandi Palassa show that the climatic division of Sri Lanka into a Wet Zone, Intermediate Zone and Dry Zone, with associated vegetation distributions of closed rainforest, intermediate tropical forest and open grassland conditions,

34

Figure 4. Fossil human and faunal stable isotope results A) δ13C and δ18O values of fauna from the Terminal Pleistocene/Holocene deposits of Fa Hien-lena, Balangoda Kuragala and Bellan-bandi Palassa, B) δ13C and δ18O values of human specimens from Late Holocene deposits Balangoda Kuragala, C) δ13C and δ18O values of human specimens from Terminal Pleistocene/ Early Holocene deposits at Fa Hien-lena and Balangoda Kuragala, D) δ13C and δ18O values of Batadomba-lena fauna dated from 20,000 to 12,000 cal. years BP and humans dated to c. 20,000 cal. years BP. Adapted from Roberts et al. (2015b). © Patrick Roberts.

35

Early Human Adaptation to Late Pleistocene-Holocene Rainforests in South Asia existed across the Terminal Pleistocene and Holocene (Roberts et al., 2015a,b) (Figure 4). Forest fauna δ13C and δ18O at Batadomba-lena indicates that Sri Lanka’s Wet Zone rainforest was more open during the Last Glacial Maximum than during the Terminal Pleistocene/Holocene. Terminal Pleistocene and Holocene human hunter-gatherer δ13C and δ18O demonstrated long-term reliance on rainforest resources (Roberts et al., 2015b) (Figure 4). A single individual from Fa Hien-lena dated to c. 12,000 cal. years BP occupied either closed Wet Zone rainforest or more open Intermediate rainforest.

The same is true of the Balangoda Kuragala human specimens dated to between 12,000 and 3,000 cal. years BP, although a distinct preference exists for Intermediate rainforest. Only two individuals, dated to c. 3,000 cal. years BP, showed any significant

contribution of open grassland resources at a time when Iron Age farming likely expanded into the region (Deraniyagala, 2007; Roberts et al., 2015b). This pattern of rainforest preference extends back to c. 20,000 cal. years BP at Batadomba-lena, when a preference for Intermediate rainforest was aided by the slight opening of the Wet Zone rainforest canopy (Roberts et al., 2015b). Stable carbon and oxygen analysis therefore demonstrates that hunter-gatherer dietary reliance on rainforest resources was therefore both possible and preferable in Late Pleistocene and Holocene Sri Lanka, even when open, grassland resources were available.

Late Pleistocene-Holocene Human Adaptation to Sri Lanka’s Rainforests: a Stable, yet Sensitive, Relationship The evidence from stable carbon and oxygen isotope analysis of fossil human and faunal tooth enamel provides direct support for technological, zooarchaeological and archaeobotanical evidence for an early and stable rainforest adaptation by Homo

sapiens in Sri Lanka. Not only do the earliest bone technologies and microlithic toolkits in Sri Lanka c. 36 Ka occur in the modern day Wet Zone, but they persist

here until c. 12 Ka and 3 Ka, respectively (Deraniyagala, 1992; Perera, 2010, Perera et al., 2011; Roberts et al., 2015a). These technological records are associated with a persistent focus on rainforest fauna and flora at the sites of Batadomba-lena and Fa Hien-lena, including particular focus on arboreal and semi-arboreal primates (Perera, 2010; Perera et al., 2011). Although detailed use-wear studies are yet to be carried out on the microlithic and bone toolkits at these sites, it is likely that they formed part of complex hunting strategies, either as projectiles or trap mechanisms (Perera et al., 2016, in press). From at least 20 Ka to 3 Ka stable carbon and oxygen isotope analysis provides direct evidence for human rainforest resource reliance in the Wet and Intermediate rainforest zones of Sri Lanka, suggesting not only Late Pleistocene rainforest use, but also Late Pleistocene rainforest specialization (Roberts et al., 2015b). This rainforest adaptation persists through periods of significant climatic and environmental change in Sri Lanka. At the time of the Last Glacial Maximum, stable carbon and oxygen isotope evidence of fossil fauna from Batadomba-lena indicates that the Wet Zone rainforests were more open at this point in time (Figure

36

4) (Roberts et al., 2015b). This fits with evidence for significant changes in tropical forest structure and extent around the globe at this time (Colinvaux et al., 1996; Dupont et al., 2000; Haberle et al., 2012). Nevertheless, Sri Lankan human foragers demonstrate persistent rainforest resource reliance during this period, albeit with a preference for more open Intermediate rainforest resources (Figure 4). Rainforest resource reliance also persists amongst human hunter-gatherers even following the introduction of agriculture to the Sri Lankan Wet Zone. Although two Iron Age individuals show a preference for ‘open’ resources c. 3 Ka, a number of human individuals retain rainforest resource reliance at this time. This not only implies

complexity in the interaction of microlithic foragers and agricultural communities, as seen on the Indian subcontinent (for example, Roy, 2008), but also demonstrates that rainforest foraging remained a stable and preferable adaptation in the Sri Lankan Wet Zone, even after agriculture entered the region. Despite the prominent and persistent place of rainforest environments in Sri Lankan prehistory, the forests of the island’s Wet Zone are now coming under threat from agricultural, industrial and urban encroachment. While the archaeological sites of Batadomba-lena, Fa Hien-lena and Kitulgala Beli-lena are now protected as a World Heritage site, this mandate does not include the protection of their modern rainforest context. The archaeology of these caves are of global importance not just because of their evidence for early Homo sapiens presence in South Asia, but

also because they document some of the earliest, sustained occupation and use of rainforest environments by our species during the Late Pleistocene. It is therefore essential that the rainforest surrounding these cave and rockshelter sites today are treated as just as much a part of Sri Lankan cultural heritage as the archaeological sequences themselves and protected accordingly. A similar situation exists for the UNESCO World Heritage site of Polonnaruwa in the Dry Zone of Sri Lanka. Here, forest clearance has taken place in order to make the archaeological ruins more visible to tourists; this endangers the tropical forests and their animal inhabitants that have formed a crucial part of Sri Lanka’s heritage since the arrival of Homo sapiens.

Conclusions and Future Research I have briefly reviewed the archaeological evidence for the Late Pleistocene occupation and utilization of Sri Lanka’s rainforests by our species. Here, at the southern tip of South Asia, where the earliest arrival of humans is, accompanied by complex microlith and bone technologies, there is evidence for the early and ongoing use of rainforest flora and fauna from c. 36 Ka to 3 Ka. Stable carbon and oxygen isotope evidence from human and faunal tooth enamel provides direct

evidence for human rainforest resource reliance from c. 20 Ka to 3 Ka, demonstrating that humans specialized in rainforest exploitation. Future research will extend this methodology to the earliest human fossils found in the Wet Zone rainforests c. 36

Ka, in order to test the longevity of this reliance. Furthermore, the application of stable isotope methodologies to other regions of the world where human rainforest relationships of global significance have been identified (including Africa, South-

37

Early Human Adaptation to Late Pleistocene-Holocene Rainforests in South Asia East Asia, Melanesia and the Americas) should provide growing insight into the spatial and temporal reliance of our species on these ecologies. The Sri Lankan case is one instance of a growing recognition that there are a number of South Asian sites suitable for UNESCO World Heritage nomination and protection (Petraglia, 2014). The evidence however, although the evidence I have described has now rightly seen three Sri Lankan cave and rockshelter sites be confirmed as a combined World Heritage entity. However, the surrounding rainforest environments of these sites have yet to be included in this nomination. Given the prominent place of Sri Lanka in global discussions of early human Late Pleistocene rainforest adaptations (Deraniyagala, 1992; Perera et al., 2011; Roberts et al., 2015a; Roberts and Petraglia, 2015), the protection of the rainforest ecology that surrounds these archaeological sequences could be considered of equal importance to Sri Lankan cultural and environmental heritage. To preserve significant ‘rainforest prehistories’ like this (Mercader, 2002b; Roberts and Petraglia, 2015), a more integrated approach to heritage listings of archaeological sites and environments will be necessary.

Acknowledgements I would like to Nuria Sanz and Robin Dennell for their invitation to take part in this volume and the associated meeting in Mexico in 2015. I would also like to extend my thanks to Michael Petraglia, Julia Lee-Thorp and Peter Mitchell for continued assistance and advice in the production of this research. I am also indebted to the Department of Archaeology, Government of Sri Lanka, including S.U. Deraniyagala, Oshan Wedage and Nimal Perera for facilitating stable isotope analysis in Sri Lanka. With thanks also to Laura Lewis for her permission to use Figure 1. Finally, I would like to thank the Natural Environmental Research Council and the Boise Fund, University of Oxford for financing my ongoing research in this area.

References Altherr, S. 2007. Going to Pot. The Neotropical Bushmeat Crisis and its Impact on

Primate Populations. Sussex, UK/Munich, Germany, Care for the Wild & Pro Wildlife.

Bailey, R. C. and Headland, T. N. 1991. The tropical rain forest: Is it a productive environment for human foragers? Human Ecology, Vol. 19, No. 2, pp. 261-285. Barker, G., Barton. H., Bird, M., Daly, P., Datan, I., Dykes, A., Farr, L., Gilbertson, D., Harrisson, B., Hunt, C., Higham, T., Kealhofer, L., Krigbaum, J., Lewis, H., McLaren, S., Paz, V., Pike, A., Piper, P., Pyatt, B., Rabett, R., Reynolds, T., Rose, J., Rushworth, G., Stephens, M., Stringer, C., Thompson, J. and Turney, C. 2007. The ‘human revolution’ in lowland tropical South-East Asia: the antiquity and behaviour of anatomically modern humans at Niah Cave (Sarawak, Borneo). Journal of Human Evolution, Vol. 52, pp. 243-261.

38

Bocherens, H., Schrenk, F., Chaimanee, Y., Kullmer, O., Morike, D., Pushkina, D. and Jaeger, J-J. 2015. Flexibility of diet and habitation in Pleistocene South Asian mammals: Implications for the fate of the giant fossil ape Gigantopithecus. Quaternary International. Cerling, T. E., Hart, J. A. and Hart, T. B. 2004. Isotope ecology in the Ituri forest. Oecologia, Vol. 138, pp. 5–12. Colinvaux, P., Oliveira, P., Moreno, E., Miller, C. and Bush, M. 1996. A long pollen record from lowland Amazonia: forest and cooling in glacial times. Science, Vol. 274, pp. 85-88.

Deraniyagala, S. U. 1992. The Prehistory of Sri Lanka: An Ecological Perspective, 2nd edn. Colombo, Sri Lanka, Department of Archaeological Survey.

Deraniyagala, S. U. 2007. The prehistory and protohistory of Sri Lanka. P. L. Prematilleke, S. Bandaranayake, S. U. Deraniyagala and R. Silva (eds), The Art and Archaeology of Sri Lanka, 1. Colombo, Sri Lanka, Central Cultural Fund, pp. 1-96.

Dupont, L., Jahns, S., Marret, F. and Ning, S. 2000. Vegetation change in equatorial West Africa: time slices for the 150 Ka. Palaeogeography, Palaeoclimatology, Palaeocology, Vol. 155, pp. 95–122.

Fortier, J. 2014. The Kings of the Forest: The Cultural Resilience of Himalayan huntergatherers. Honolulu, USA, University of Hawaii Press.

Gamble, C. 1993. Timewalkers: The Prehistory of Global Colonization. Stroud, UK, Alan Sutton.

Haberle, S. G., Lentfer, C., O’Donnell, S. and Denham, T. 2012. The palaeoenvironments of Kuk Swamp from the beginnings of agriculture in the highlands of Papua New Guinea. Quaternary International, Vol. 249, pp. 129-139. Hayes, J. 1983. Practice and principles of isotopic measurements in organic geochemistry.W. Meinschein (ed.), Organic geochemistry of contemporaneous and

ancient sediments. Bloomington, Ind., Society of Economic Paleontologists and Mineralologists. pp. 5-31.

Hoefs, J. 1997. Stable Isotope Geochemistry, 4th edn. Berlin, Springer-Verlag. Koch, P. L. 2007. Isotopic study of the biology of modern and fossil vertebrates. B. Michener and K. Lajtha (eds), Stable Isotopes in Ecological and Environmental Science. Boston, Mass., Blackwell, pp. 99-154.

39

Early Human Adaptation to Late Pleistocene-Holocene Rainforests in South Asia Krigbaum, J. 2003. Neolithic subsistence patterns in northern Borneo reconstructed with stable carbon isotopes of enamel. Journal of Anthropological Archaeology, Vol. 22, pp. 292–304.

Krigbaum, J. 2005. Reconstructing human subsistence in the west mouth (Niah Cave, Sarawak) burial series using stable isotopes of carbon. Asian Perspectives, Vol. 44, pp. 73-89.

Lee-Thorp, J. A. 2008. On isotopes and old bones. Archaeometry, Vol. 50, No. 6, pp. 925-950.

Levin, N. E., Simpson, S. W., Quade, J., Cerling, T. E. and Frost, S. R. 2008. Herbivore enamel carbon isotopic composition and the environmental context of Ardipithecus at Gona, Ethiopia. J. Quade and J. G. Wynn (eds), The Geology of Early Humans in the Horn of Africa. Geological Society of America Special Paper, Vol. 446. Boulder, Colo., pp. 215-234.

Lewis, L., Perera, N. and Petraglia, M. 2014. First technological comparison of Southern African Howiesons Poort and South Asian microlithic industries: an exploration of inter-regional variability in microlithic assemblages. Quaternary International, Vol. 350, pp. 7-25.

Lu, F. 2010. Patterns of indigenous resilience in the Amazon: a case study of Huaorani hunting in Ecuador. Journal of Ecological Anthropology, Vol. 14, No. 1, pp. 5-21. Mercader, J. 2002a. Forest people: The role of African rainforests in human evolution and dispersal. Evolutionary Anthropology, Vol. 11, pp. 117-124. Mercader, J (ed.). 2002b. Under the Canopy: The archaeology of tropical rainforests. Brunswick, NJ, Rutgers University Press.

Nelson, S. 2003. The extinction of Sivapithecus: faunal and environmental changes in the Siwaliks of Pakistan. American School of Prehistoric Research Monographs, volume 1. Boston: Brill Academic Publishers. Nelson, S. 2007. Isotopic reconstructions of habitation change surrounding the extinction of Sivapithecus, a Miocene hominoid, in the Siwalik Group of Pakistan. Palaeogeography, Palaeoclimatology, Palaeoecology 243: 204-222. Papworth S, Milner-Gulland E. J. and Slocombe, K. 2013. Hunted woolly monkeys (Lagothrix poeppigii) show threat-sensitive responses to human presence. PLoS ONE, Vol. 8, No. 4.

Perera, N. 2010. Prehistoric Sri Lanka: Late Pleistocene Rockshelters and an Open Air Site BAR International Series. Oxford, UK, Archaeopress.

40

Perera, N., Kourampas, N., Simpson, I. A., Deraniyagala, S. U., Bulbeck, D., Kamminga, J., Perera, J., Fuller, D. Q., Szabo, K. and Oliveira, N. V. 2011. People of the ancient rainforest: Late Pleistocene foragers at the Batadomba-lena rockshelter, Sri Lanka. Journal of Human Evolution, Vol. 61, pp. 254-269. Perera, N., Roberts, P. and Petraglia, M. 2016. Bone technology in South Asia from Late Pleistocene rockshelter deposits in Sri Lanka.  M. C. Langley (ed.), Forthcoming.  Osseous Projectile Weaponry: Towards an Understanding of Pleistocene Cultural Variability. New York, Springer-Verlag. (VERT Series.)

Petraglia, M. 2014. Out of Africa and the evolution of human populations in Asia: thoughts about the nomination of prehistoric sites to the World Heritage List. N. Sanz (ed.), Human Origin Sites and the World Heritage Convention in Asia. Mexico City, United Nations Educational, Scientific and Cultural Organization.

Roberts, P. and Petraglia, M. 2015. Pleistocene rainforests: barriers or attractive environments for early human foragers. World Archaeology, Vol. 47, No. 5, pp. 718-739.

Roberts, P., Boivin, N. and Petraglia, M. D. 2015a. The Sri Lankan ‘microlithic’ tradition c. 38,000 to 3000 years ago: tropical technologies and adaptations of Homo sapiens at the southern edge of Asia. Journal of World Prehistory, Vol. 28, No. 69.

Roberts, P., Perera, N., Wedage, O., Deraniyagala, S., Perera, J., Eregama, S., Gledhill, A., Petraglia, M. D. and Lee-Thorp, J. A. 2015b. Direct evidence for human reliance on rainforest resources in late Pleistocene Sri Lanka. Science, Vol. 347, pp. 1246-1249.

Roy, B. 2008. Modern microlith makers in Mandla, Madhya Pradesh (India): continuity or re-invention? Antiquity, Vol. 82, No. 31. Sharp, Z. 2006. Principles of stable isotope geochemistry. Upper Saddle River, NJ, Pearson Prentice Hall.

Somadeva, R. and Ranasinghe, R. 2006. An excavation of a shell midden at Pallemalla in southern littoral area of Sri Lanka. Ancient Asia, Vol. 1, pp. 14-24. Sponheimer, M., Alemseged, Z., Cerling, T. E., Grine, F. E., Kibel, W. H., Leakey, M. G., Lee-Thorp, J. A., Manthi F. K., Reed, K. E., Wood, B. A. and Wynn, J. G. 2013. Isotopic evidence of early hominin diets. Proceedings of the National Academy of Sciences of the United States of America, Vol. 110, No. 26, pp. 10513-10518.

Summerhayes, G. R., Leavesley, M., Fairbairn, A., Mandui, H., Field, J., Ford, A., Fullagar, R. 2010. Human adaptation and plant use in Highland New Guinea 49,000 to 44,000 years ago. Science, Vol. 330, pp. 78-81.

41

Sponheimer, M., Alemseged, Z., Cerling, T. E., Grine, F. E., Kibel, W. H., Leakey, M. G., Lee-Thorp, J. A., Manthi F. K., Reed, K. E., Wood, B. A. and Wynn, J. G. 2013. Isotopic evidence of early hominin diets. Proceedings of the National Academy of Sciences of the United States of America, Vol. 110, No. 26, pp. 10513-10518.

Van der Merwe, N. J. and Medina, E. 1991. The canopy effect, carbon isotope ratios and foodwebs in Amazonia. Journal of Archaeological Science, Vol. 18, pp. 249-259.

White, T. D., Asfaw, B., Beyene, Y., Haile-Selassie, Y., Owen Lovejoy, C., Suwa, G., WoldeGabriel, G. 2009. Ardipithecus ramidus and the Paleobiology of Early Hominids. Science, Vol. 326, No. 5949, pp. 75-86. Wijeyapala, W. H. 1997. New light on the prehistory of Sri Lanka in the context of recent investigations of cave sites. Ph.D. thesis, University of Peradeniya, Sri Lanka.

42

Stéphen Rostain Le Centre national de la recherche scientifique (CNRS), Francia

Abstract

Resumen

The Amazon is the world’s largest rainforest and is

La Amazonia es la selva tropical más grande del mun-

characterized by an exceptional biodiversity. The

do y se caracteriza por una biodiversidad excepcio-

assessment of the pre-Columbian landscape of this

nal. La evaluación del paisaje precolombino de esta

region has experienced a quantum leap forward

región ha dado un salto prodigioso desde que la ar-

since archaeology adopted the historical ecology

queología adoptó la perspectiva de la ecología histó-

perspective. Let’s not forget that, until a few years

rica. Hay que recordar que hasta hace pocos años, los

ago, researchers saw the Amazon as a primary tropical

investigadores consideraban a la Amazonia como un

rainforest, unaltered by indigenous peoples. The

bosque húmedo tropical primario, no modificado por

reality is much more complex.

las poblaciones indígenas. La realidad es mucho mas compleja que ello.

Recent scientific research shows great human influence in this immense territory during the pre-Columbian

Las investigaciones científicas recientes muestran una

era. Man intervened in the composition of flora and

gran influencia humana durante la época precolom-

distribution of vegetation, as well as in the creation of

bina en este inmenso territorio. El hombre intervino

highly fertile black soils.

en la composición florística y en la distribución de la vegetación, así como en la creación de suelos negros

Great works, of different types, achieved by the first

de alta fertilidad.

inhabitants persist today in the Amazonian landscape. We can mention seven sites from this pre-Columbian

Diferentes tipos de grandes obras realizadas por los pri-

inheritance that deserve particular attention and

meros habitantes quedan hoy en el paisaje amazónico.

protection: raised fields (ridges); residential or

De esta herencia precolombina, podemos mencionar

funeral earth mounds; sambaquis (shell middens); pits,

siete  tipos de sitios espectaculares que merecen una

megalithic sites; and rock art sites.

dos (camellones); los montículos de tierra residenciales

sometimes called geoglyphs; terra preta (dark soil);

atención particular y ser protegidas: los campos elevao funerarios; los sambaquis (conchales); las fosas, a veces

llamadas geóglifos; la terra preta (tierra oscura); los sitios megalíticos, y los sitios de arte rupestre.

44

Sitios monumentales precolombinos en Amazonia Amazonia, la selva tropical más grande del mundo, se caracteriza por una biodiversidad excepcional. En paralelo, también existe una importante diversidad cultural, misma que cumplió un papel esencial en la constitución del paisaje actual. La evaluación del paisaje precolombino de esta región ha visto un salto prodigioso desde que la arqueología adoptó la perspectiva de la ecología histórica (Balée & Erickson, 2006). Hay que recordar que todavía hasta hace pocos años, los investigadores consideraban a la Amazonia como un bosque primario tropical húmedo no modificado por las poblaciones indígenas. Sin embargo, la realidad se presenta aún más compleja que esto. Las recientes investigaciones científicas muestran una gran influencia humana en el territorio durante la época precolombina, con intervención del hombre en la composición florística, la distribución de la vegetación, así como también en la creación de suelos negros de alta fertilidad. Asimismo, al cavar y levantar el suelo, éste jugó a veces un rol preponderante en la formación del modelado, esculpiendo la superficie del bosque. Su impacto ha marcado entonces a la Amazonia desde hace tiempo. El medio ambiente actual de este bosque es en parte resultado de su acción secular. Hace 20 años, el etnólogo Claude Lévi-Strauss adoptó la idea de una Amazonia cultural precolombina mucho más diversa y compleja, misma que fue revelada paulatinamente por los arqueólogos de aquella época. Antes, se consideraba que “la ilusión de que la condición actual de las comunidades indígenas prolongaba aquella que era la suya en el momento del descrubrimiento, estaba firmemente instalada”, pero, en realidad, “la selva amazónica no es en todas partes tan ‘primaria’ como se solía pretender. [...] La Amazonía podría ser la cuna de donde surgieron las civilizaciones andinas” (Lévi-Strauss, 1994, pp. 12-13). En efecto, la nueva generación de arqueólogos que trabajan en esta selva desde hace 30 años ha ido cambiando de manera radical la opinión general sobre los pueblos precolombinos, sacando a la luz la existencia de grandes asentamientos en varios lugares, así como también de comunidades complejas importantes que no llegaron por medio de una migración lejana originaria de los Andes, sino que son resultado de un largo desarrollo local.

45

Sitios monumentales precolombinos en Amazonia

Figura 1. Mapa de los sitios monumentales precolombinos de la Amazonia citados en el texto. © S. Rostain.

Grandes obras realizadas por los primeros habitantes se observan hoy en el paisaje amazónico (Rostain, 2011; Walker, 2012). De esta herencia precolombina, podemos retener siete tipos de sitios merecedores de una atención particular y con miras a ser protegidos: los campos elevados, los terraplenes, las fosas, los montículos de conchas, la terra preta, los megalitos y el arte rupestre (Figura 1).

El mosaico natural y cultural de la Amazonia Hablar de la Amazonia es a menudo subrayar su extrema biodiversidad, tema que ha llamado la atención de científicos y del mundo. Hay que precisar que las cifras son impresionantes: • más de 16,000 especies de árboles diferentes (cuando hay solamente 140 en Francia), de las cuales 5,000 han sido formalmente identificadas; • una hectárea de bosque denso cuenta con 200 a 400 especies de árboles frente a una veintena en un bosque temperado; • más de 40,000 especies de plantas; • 2.5 millones de especies de insectos; • 15% de las especies vegetales y animales del mundo. Pero es necesario comprender bien lo que se lee detrás de estas cifras. Por ejemplo, si son 16 mil las especies de árboles diferentes estimadas en la Amazonia, solamente 227 representan la mitad de los árboles. Entre éstas, las tres más corrientes son la palmera Euterpe precatoria y los árboles Protium altissimum y Eschweilera coriácea. Estas

46

tres especies equivalen a 4% de todo el bosque, es decir, a 5 millares de troncos de cada una (Steege et al., 2013). Entonces con un promedio de 500 árboles por hectárea, la Amazonia es menos densa e impenetrable de lo que uno se imagina dado que su densidad es comparable con aquella de los bosques franceses. Sin embargo, debido a esta atención excepcional prestada a la fauna y a la flora, la diversidad humana ha sido por desgracia ignorada. Se vuelve ilusorio reducir la antigua ocupación de la Amazonia a un solo modelo cuando en este territorio del tamaño de Europa las formas de sociedad fueron tan diversas como los tipos de entorno (Rostain, 2016). Allí, los cazadores-recolectores nómadas coexistieron con grandes comunidades edificadoras de sitios majestuosos. Es esta pluralidad la que el arqueólogo comienza apenas hoy a revelar. Además, los amerindios fueron quienes transformaron profundamente este bosque. En su capa vegetal, en la naturaleza de sus sedimentos o incluso en el modelado del suelo, estos cambios, voluntarios o no, fueron observados en primer lugar por los geógrafos, y los antropólogos o los arqueólogos lo vienen haciendo desde hace ya unas tres décadas. El manto verde que cubre la región es menos natural de lo que su exuberancia deja pensar. El hombre deshierbó sembró, multiplicó, cruzó, asoció o mejoró las especies. William Balée (1989) estima que la vegetación de alrededor de 12% de la cuenca amazónica fue manipulada por los indígenas. La minuciosa observación de ésta no deja ninguna duda sobre la importancia que tuvo la intervención humana antigua en su estado actual. De manera comparable, los estudiosos sacaron a la luz suelos parcialmente creados por el hombre. Las terras pretas son suelos compuestos, oscuros y fértiles asociados a restos de implantaciones y enriquecidos con desechos de ocupación, carbón y cenizas. Muchos investigadores se han preguntado si son resultado de ocupaciones largas o sucesivas en un mismo emplazamiento y si son o no fruto de actos creativos voluntarios. De cualquier modo queda claro que son islotes de fertilidad sorprendente en un universo pedológico ácido y pobre. Se vuelve difícil creer que esta notable cualidad no haya sido observada por los primeros habitantes de la Amazonia, aunque todavía sea complicado probar que los mismos suelos se usaron anteriormente para la agricultura. Si pudiéramos responder afirmativamente a esta pregunta confirmaríamos un modelo clave de ecología histórica, ya que la ocupación de localidades durante una cierta época precolombina ayudó a amplificar su capacidad productiva. Finalmente, no sólo la capa vegetal y la naturaleza del suelo fueron modificadas por el hombre, sino también la morfología de la región. El amerindio cavó, volteó, transportó, levantó y edificó miles de metros cúbicos de tierra, transformando más o menos de forma radical el modelado de su territorio. Debemos imaginar una Amazonia precolombina atravesada por rutas permanentes, canales y fosas entrecruzadas, caminos elevados que conectan lomas y montículos, diques que contienen bahías, reservorios o campos elevados de todas las formas, dimensiones y arreglos posibles.

47

Sitios monumentales precolombinos en Amazonia Así, la Amazonia conserva tanto en su suelo, su superficie como vegetación, la huella discreta de sus primeros habitantes. Estas son las marcas antrópicas que vamos a detallar a continuación.

Campos elevados Hoy, la mayor parte de los grupos amerindios utiliza la técnica de quema-y-rosa, la cual consiste fundamentalmente en abrir una parcela en el bosque con tala de árboles y quema. Las cenizas sirven para fertilizar los suelos pobres y los campos son cultivados algunos años antes del desplazamiento. Existen numerosas variaciones de esta técnica, como la tala y quema o la agroforestal, pero en la época precolombina existieron además otras técnicas monumentales ya olvidadas en la actualidad (Denevan, 2001); entre ellas, las más espectaculares son los campos elevados. Éstos son montículos destinados a la agricultura y están presentes en numerosos países de América Latina. Varios sitios de campos elevados se hallan situados en las alturas de los Andes. En su mayoría construidos durante el primer milenio de nuestra era, algunos de ellos se remontan no obstante al año 1000 a. C. Allí, a la función de drenaje se suma la utilidad térmica, pues la presencia de mucha agua equilibra las variaciones de temperatura diurnas/nocturnas. Por otra parte, aunque ciertos campos elevados fueron construidos en áreas inundables, muchos se hallan más bien al borde de un lago y no así de un río, en donde las necesidades de irrigación completan aquellas de drenaje. En las tierras bajas, los sitios de campos elevados son esencialmente camellones (montículos alargados) y sólo a veces lomas redondas o cuadrangulares. Existen campos elevados precolombinos en las Guayanas, Venezuela, Colombia, Ecuador y Bolivia. Éstos, en los llanos de Mojos, en Bolivia, se extienden a lo largo de una superficie de 6,000 hectáreas y son de formas muy variadas (Figura 2). En esta inmensa región, pueden diferenciarse seis tipos de paisajes agrícolas precolombinos (Walker, 2012).

Figura 2. Grandes campos elevados de los llanos de Mojos, Bolivia. © 2017 Google, Digital Globe.

48

Figura 3. Campos elevados del sitio Diamant, Guayana Francesa. © S. Rostain.

En las Guayanas, es decir en la Amazonia atlántica, las poblaciones de tradición arauquinoide (650-1400 d. C.) ocuparon un territorio litoral de 600 km de largo en el cual la técnica de campos elevados fue ampliamente utilizada durante cerca de un milenio, antes de la llegada de los europeos (Rostain, 2012a). Estos campos se hallan entre el río Berbice, en Guyana oriental, y la Isla de Cayena, en Guayana Francesa (Figura 3). Los primeros campos elevados arauquinoides se construyeron alrededor del 650 d. C. en el oeste del Surinam, extendiéndose por toda la costa guayanense a partir de 1000 d. C. y hasta 1400 d. C. Su densidad más alta se encuentra al este del territorio arauquinoide, en Guayana Francesa, lugar en el que habitaron las últimas comunidades de esta tradición. Siete diferentes sectores yuxtapuestos fueron reconocidos a lo largo de la costa donde se encuentran los campos elevados. Los complejos presentan una organización específica dominante en cada uno de estos sectores. La naturaleza y la distribución de los montículos agrícolas están ligadas a la topografía y al nivel del agua, pero pueden existir variaciones de formas de organización al interior de un mismo com

49

Sitios monumentales precolombinos en Amazonia plejo. La localización topográfica es indicativa de diferentes adaptaciones a las condiciones hidrográficas y a la naturaleza del suelo. Los campos elevados grandes se hallan en las áreas anegadas y los de tamaño medio siguen las curvas de los talwegs,

la mayoría dispuestos a lo largo de pendientes. Los pequeños y medianos, en cambio, cubren grandes sabanas inundables. Se conocen ahora datos sobre la historia del paisaje en donde se encuentran los campos elevados. Una muestra de sedimento que representa 2,150 años de evolución vegetal fue tomada en una ciénega delante de un gran complejo de éstos (Iriarte et al. 2012). Los resultados combinan los análisis de polen, de fitolitos y de carbón y revelan una ausencia de incendios de sabana durante el cultivo de los montículos en el período precolombino. Un fuerte aumento de los fuegos es observable tras su abandono, durante la llegada de los europeos. Al contrario de lo que se afirmaba hasta entonces, se descubre que los amerindios controlaban y limitaban los fuegos para mejorar la producción agrícola. El hombre impuso entonces regularidad, simetría, geometría y recurrencia en la agricultura. Esto se aplica de igual manera en el caso de la Amazonia, incluso si la lógica utilizada no es siempre evidente a primera vista. Los amerindios clasificaron y transformaron su territorio siguiendo una verdadera ciencia.

Terraplenes Los antiguos habitantes de la Amazonia fueron ante todo constructores únicos que cavaron y amontonaron tierra sin descanso para así edificar montículos, caminos, diques, canales o reservorios (Rostain, 2011). De todos estos monumentos, la atención se ha centrado de forma prioritaria en los montículos grandes, dejando de lado construcciones menos espectaculares. No obstante, es necesario a menudo considerar todo el conjunto del paisaje transformado para comprender la real intención de la obra. Los imponentes montículos de hábitat están frecuentemente acompañados por caminos o murallas elevadas y a los montículos secundarios se han añadido fosas, desembarcaderos, canales o pequeños charcos artificiales. Aunque estos conjuntos tienen a veces modestas dimensiones, como las elevaciones domésticas bajas del medio Amazonas (Neves, 2012), se conocen también montículos altos, como aquellos de los llanos de Venezuela, aquellos a lo largo de la costa de las Guayanas, en la desembocadura del Amazonas y en el costado opuesto, en la Amazonia occidental; en los llanos de Mojos, en Bolivia, y a lo largo del piedemonte de los Andes, en Ecuador. En la mayor parte de los casos, otras estructuras están asociadas a estos montículos. Las excavaciones precolombinas de la Amazonia son de diferentes tipos. Se pueden identificar tres funciones principales en ellas. La primera función, señalada anteriormente, es la agricultura en campos elevados. La segunda es la habitacional en su cima, uso al cual se puede sumar la función funeraria y a veces también una utilización ceremonial, puesto que las sepulturas estaban frecuentemente situadas en estos

50

montículos residenciales. Este grupo concierne a las fosas periféricas que parecen tener una función defensiva, ritual o astronómica. Las obras más antiguas conocidas en la Amazonia están ubicadas en el valle del Upano, en Ecuador, al pie de los Andes (Rostain, 2012b). Hace 2,500 años, varios grupos humanos comenzaron a asentarse a lo largo del barranco que domina el río Upano. Allí construyeron plataformas, plazas planas, bajas y caminos cavados (Figura 4). En varios casos, aprovecharon la pendiente o un relieve natural del lugar para erigir el montículo, con lo que ahorraron trabajo de pala o transporte de terraplén. Debido al sedimento lodoso y resbaloso en temporada de lluvias, éstos quemaron la cima de los montículos para obtener un suelo compacto y duro. Los montículos tienen formas ovaladas o rectangulares, en algunos casos en L o en T, y excepcionalmente circulares. La superficie superior es plana o ligeramente convexa y mide en su cima entre 10 y 50 m de largo, y de 3 a 10 m de ancho, con una altura variable de 2 a 10 m. En el sitio de Sangay, el más importante de la región, tres caminos cavados largos, rectilíneos y paralelos atraviesan el espacio, siendo el más importante aquel que mide 1,200 m de largo, con un ancho que puede alcanzar los 13 m y cuya profundidad actual es de 3 m, pues ha sido parcialmente rellenado por desmoronamientos de los Figura 4. Complejo de montículos artificiales de tierra del sitio Edén, valle del Upano, Ecuador. © S. Rostain.

bordes. Éstos se hallan conectados perpendicularmente por caminos secundarios y provistos de un reborde en uno o sus dos costados, y se extienden a lo largo de dece

51

Sitios monumentales precolombinos en Amazonia nas o centenas de metros con un ancho en su base de 2.5 a 10 m y una profundidad superior a 3 m. La característica excepcional de estos sitios es su organización con un modelo espacial recurrente que consiste en una plaza central cerrada por montículos. Éste tiene diversas variaciones, como por ejemplo, la presencia de un montículo central o seis periféricos. Hay caminos cavados que suavizan la pendiente natural hasta el riachuelo vecino y que están a menudo conectados a los complejos. La primera función de la mayor parte de las elevaciones fue la residencial; sin embargo existen algunos ejemplos de montículos periféricos particularmente estrechos que debieron cumplir con otro uso aún indefinido. La disposición particular de los complejos de montículos y plazas permite pensar en una posible función ceremonial secundaria. Son cientos de complejos los que fueron encontrados. En realidad, las recientes prospecciones aéreas con fotografías Lidar han revelado una extensión espectacular de estas obras que ocupan prácticamente toda la superficie del valle. La cuenca entera fue transformada y acondicionada por los habitantes precolombinos. Además, el hecho de que muchos de estos sitios estén conectados por caminos sugiere que fueron contemporáneos. Más al sur, en los llanos de Mojos, en Bolivia, se encuentran muchas obras de tierra precolombinas. Son canales, reservorios, taludes de trampa para peces, caminos elevados, montículos residenciales y campos elevados (Erickson, 2008). Los montículos conocidos localmente como “lomas” se hallan casi siempre agrupados y están distribuidos a lo largo de un curso de agua o alrededor de un espacio plano o un charco de agua. Son ovalados o redondos, algunos en arco de círculo a orillas de un espacio inundado, existen dobles y hasta múltiples. Los vestigios hallados en la mayoría de los montículos dan muestra de su uso doméstico pero igualmente funerario. Los montículos grandes alcanzan de 3 a 8 m de altura y su número estaría entre 200 y 300. Habría varios miles de medianos de 1 a 3 m de altura que sirvieron para casas. Finalmente, están aquellos con una función particular, de forma larga o rectangular, cuya disposición paralela alrededor de una plaza y la localización alejada del hábitat conducen a pensar en su uso ceremonial. A decir verdad, las lomas son una de las numerosas construcciones humanas de los llanos de Mojos, cuyo paisaje fue profundamente transformado por el hombre. Por ejemplo, el montículo de la Loma Salvatierra, de 8 m de altura y ocupado entre el 500 y 1400 d. C., está rodeado por un talud periférico que circunscribe un área de 20 ha a su alrededor (Lombardo & Prümers, 2010). El hábitat está separado del cementerio y el entorno organizado por una red de canales, fosas, reservorios circulares, pequeños montículos y caminos elevados.

52

En la costa de las Guayanas, entre los ríos Berbice en Guyana y Coppename en Surinam, no hay elevaciones naturales. Así, los habitantes se vieron obligados a edificar montículos de arcilla para asentar sus pueblos sobre el nivel del agua. Estaban ubicados en el punto de encuentro de las aguas dulces y las aguas saladas, facilitando la explotación de los recursos marinos y terrestres ubicados a poca distancia del pueblo. Los alrededores de los montículos se inundan de agua dulce al fin de la temporada seca. Varias líneas naturales de agua fueron acondicionadas por los indígenas o, a veces, creadas en su totalidad. Hertenrits, construido a partir del 650 d. C., es el montículo más grande, cuyo diámetro abarca de 200 a 320 m y cuya altura alcanza los 2.5 m (Figura 5). Está rodeado por una depresión de 20 a 100 m de ancho y su edificación fue hecha por medio de extracción de arcilla. Cinco desembarcaderos se acondicionaron en su perímetro para facilitar el atracamiento de canoas que circulaban en los canales. Dos montículos satélites más pequeños fueron edificados diametralmente opuestos a equidistancia de Hertenrits. En los alrededores, hay grupos irregulares y dispersos de campos elevados. Varios canales estrechos e inundados anualmente tienen una

Figura 5. Entornos acondicionados del montículo artificial de Hertenrits, Surinam, con canales, caminos cavados y campos elevados. © S. Rostain.

53

Sitios monumentales precolombinos en Amazonia disposición radial desde Hertenrits y llegan a los campos elevados y montículos satélites. Sirven de camino durante la temporada seca y para circular en canoa durante la temporada húmeda, lo que comprueba que los montículos estuvieron ocupados en la misma época. En la desembocadura del Amazonas en Brasil, en la Isla de Marajó, se encontraron igualmente complejos de montículos de tierra. Allí éstos tienen una altura promedio de 3 a 7 m, pero pueden alcanzar hasta 20 m y miden de 600 a 2,500 m2; algunos se extienden a veces cerca de 900,000 m2. Fueron edificados por las comunidades Marajoara entre el 450 y 1350 d. C. (Roosevelt, 1991). Los montículos residenciales se distinguen de los ceremoniales, aparentemente reservados a una élite. En estos últimos se hallaron enterradas urnas funerarias polícromas (Schaan, 2011). Están reunidos conformando grandes implantaciones. Si todos los montículos de Camutins, el sitio más importante, estuvieron ocupados a la vez, esto representaría entre 28 y 43 grandes casas comunitarias, es decir, cerca de 2,400 habitantes (Roosevelt, 1991). A fin de constituir reservas durante la temporada seca, cerca de los montículos se acondicionaron, al parecer, piscinas de peces. La construcción de retenciones de agua en los ríos habría permitido una pesca abundante que proveyó de suficientes proteínas, facilitando así el crecimiento demográfico. Se ha sugerido incluso que las sociedades Marajoara dependían esencialmente de la explotación intensiva de los recursos acuáticos, dispensándose así la intensificación de la producción agrícola para cubrir las necesidades de una población numerosa (Schaan, 2011). La abundancia de vestigios de actividades ceremoniales y funerarias en los montículos mayores sugiere su estatus de centros políticos y ceremoniales a los cuales se subordinaron los sitios más pequeños. A esta cultura Marajoara, al igual que a otras contemporáneas de tradiciones Polícroma o Arauquinoide en la Amazonia, se les considera como aquellas que marcaron la emergencia de una complejidad social en las tierras bajas amazónicas. Resultado del choque microbiano a causa de la conquista europea y de la desestructuración sociopolítica que con ella sobrevino, las sociedades amazónicas actuales conservaron poco de esta dinámica de terraplenes. Sin embargo, todavía podemos identificar huellas de este pasado en algunos grupos. Por ejemplo, las poblaciones Kayapó del Alto Xingú, en Brasil, viven aún en grandes pueblos anulares conectados entre ellos por rutas muy anchas y rectilíneas. Toda la región está de este modo organizada por una red de caminos y pueblos. Esta distribución espacial es en realidad herencia de un sistema más importante de antes de la llegada de los europeos. Los estudios arqueológicos de Michael Heckenberger (2005) revelaron la existencia de un antiguo territorio organizado por centros principales e implantaciones secundarias que atestiguan la existencia de una sociedad estratificada. Así, el sitio de Kuhikugu, el mismo que conociera su mayor

54

desarrollo hacia 1510 d. C., era una implantación anular de 50 ha provista de una inmensa plaza central con grandes casas comunales en la periferia y ceñida por dos fosas de talud con vocación defensiva. Desde el pueblo arrancaban radialmente rutas muy anchas hasta los sitios secundarios. Se realizaron diversos terraplenados mayores o menores en el paisaje contiguo: desembarcaderos, cimientos de puentes, etcétera.

Fosas En algunos casos, los terraplenes consisten solamente en fosas, pero pueden también alcanzar proporciones enormes. Los sitios monumentales con fosas más impactantes se encuentran en el estado de Baures, en Bolivia, y, en especial, en el estado de Acre, en Brasil, ocupando una superficie de más de 50,000 km2 en el alto Purus. Fueron hallados gracias a la intensificación de la deforestación desde hace 40 años, la misma que ha dejado amplios espacios sin árboles en los cuales aparecen claramente las estructuras cavadas. Son cientos de sitios con fosa periférica redonda, ovalada o cuadrangular, de 90 a 320 m de diámetro (Figura 6). Hay que señalar que los círculos, cuadrados y figuras geométricas son perfectamente regulares, lo que indica el uso de un instrumento de medida muy preciso (Schaan, 2011).

Figura 6. Fosas geométricas del sitio Fazenda Colorada, estado de Acre, Brasil. © M. Pärssinen.

55

Sitios monumentales precolombinos en Amazonia

Figura 7. Fosa periférica del sitio Granja del Padre, llanos de Mojos, Bolivia. © H. Prümers/DAI.

Fueron construidos y ocupados de manera discontinua del año 0 al 1300 d. C. La función de estos sitios es difícil de establecer, pues si la construcción de estas grandes fosas implica una población importante, pocos vestigios culturales, correspondientes más bien a pequeños grupos, fueron encontrados durante las excavaciones arqueológicas. El uso más probable es el ritual, tal vez como lugares de reunión para ceremonias (Schaan, 2014). Otra posibilidad es un variedad de uso según la época y la ubicación (Prümers, 2014). Así, en el norte de Bolivia, los sitios con zanja periférica, fechados entre 1200 d. C. y la época colonial, fueron considerados para tener una función defensiva (Figura 7). En el costado opuesto de la Amazonia, en las Guayanas, existe una multitud de “montañas coronadas”, nombre local dado a las elevaciones con fosas periféricas en su cima. Se trata generalmente de pequeñas colinas cercanas a un curso de agua, ceñidas en su cima por una fosa de 3 a 15 m de ancho por 1 m de profundidad, a menudo rellenado en uno, dos o tres lugares, con el fin de acondicionar el paso. La fosa mide en promedio un centenar de metros de diámetro. Algunas alcanzan los 800 m de largo, 8 m de ancho y 5 m de profundidad. En uno de los casos, la fosa fue directamente cavada en la roca. A veces, sobre todo en las orillas altas del río, la fosa se limita a un espolón cerrado. Diferentes colinas con fosa de Guayana Francesa dieron fechas entre el inicio de nuestra era y la conquista europea. La tradición oral de los Wayampi del Oyapock menciona estos antiguos sitios karanes, nación enemiga que habitaba en pueblos con estructura defensiva: “en toda la región había muchos Kalana. Construían pueblos que protegían con fosas de dos metros de ancho por uno de profundidad. La base estaba plantada de estacas. Aque-

56

Figura 8. Montículo de conchas de Taperinha, abajo de Santarém, Pará, Brasil. © A. C. Roosevelt.

llos que no lo sabían, al caer en ellas morían. Existen todavía restos de estas fosas alrededor de una montaña cerca de Camopi y en otros lugares” (Capitan Norbert citado por Grenand, 1982, p. 270). Sabemos, además, según los archivos, que los pueblos Wayana de Guayana Francesa presentaban sistemas defensivos elaborados en Guayana a fines del siglo XVIII. Podemos entonces pensar razonablemente que las colinas con fosa tenían una vocación defensiva.

Sambaquis Los precolombinos no se contentaron con utilizar la tierra o la arcilla como material de construcción, sino que hallaron además una materia prima rara y original. Desde la antigüedad recuperaron toneladas de conchas en las colonias de éstas, ya muertas y secas, para así edificar montículos de dimensiones impresionantes. Alzadas progresivamente a lo largo de los siglos, las estructuras terminaron alcanzando alturas considerables. Se conocen estos amontonamientos sobre todo en la costa de Brasil, en donde se los denomina sambaquis, pero están igualmente presentes en la costa atlántica de la Amazonia y en grandes ríos del interior. Aparte de sus gigantescas

proporciones, estos sitios proporcionaron revelaciones sorprendentes sobre los primeros habitantes de esas orillas. El sambaqui amazónico más célebre es ciertamente el de Taperinha, cerca de Santarém, en el bajo Amazonas, en Brasil (Figura 8). Allí, en el abrigo rocoso de Pedra Pintada, se descubrieron las cerámicas más antiguas del continente americano, mismas que habrían aparecido hacia 7,080 años A. P. en Taperinha, e incluso más temprano en Pedra Pintada, hacia 7,600 años A. P. (Roosevelt et al., 1991). Taper-

inha, con una altura de 6 m, se extiende a lo largo de varias hectáreas y fue ocupado durante cerca de 1,300 años. Su cerámica consiste de forma esencial en cuencos simples ocasionalmente decorados con algunas incisiones. Más hacia el oeste, a lo largo del Guaporé, se halla otro amontonamiento conchal aislado (Figura 9). Este sambaqui de Monte Castelo alcanza unos 8 m de alto y de 110 a 150 m de diámetro. Todavía se observa en él una larga secuencia de ocupa-

57

Sitios monumentales precolombinos en Amazonia

Figura 9. Sambaqui de Monte Castelo, alto Guaporé, Brasil. © E. Neves. ción que comenzó tal vez hace 8 mil años (Neves, com. pers.). Las excavaciones arqueológicas dieron como resultado un rico conjunto de sepulturas funerarias, terra preta y diferentes rasgos, como también vestigios muy variados.

En dos porciones de la Amazonia litoral, se han identificado igualmente sambaquis.

Pocas regiones ofrecen recursos importantes en conchas a lo largo de la costa de Guyana. Éstas se hallan más bien circunscritas al flujo cenagoso transcostero del Amazonas, que va desde su desembocadura hacia el noroeste, hasta la costa central de Guyana. Estos bancos cenagosos impidieron la proliferación de conchas en las costas de Amapá, Guayana Francesa, Surinam y el este de Guyana. En cambio, en la costa del Pará al este del Amazonas, en el litoral de Guyana al oeste del río Essequibo, existían grandes colonias de conchas. Los pescadores recolectores colonizaron estas dos regiones ya que allí encontraron grandes cantidades de conchas que utilizaron para edificar sus montículos. Su interés por este tipo de restos como material de construcción los condujo hasta estos lugares. La cultura Alaka ocupaba la costa oeste de Guyana, mientras que los grupos Mina vivían en el litoral del Pará en Brasil. Los vestigios de sus pueblos sobre los amontonamientos de concha representan una de las huellas arqueológicas más impresionantes de la región. Las implantaciones Alaka son las más antiguas, con fechas de entre 6,800 y 3,550 años A. P., y los establecimientos Mina tienen fechas entre 5,700 y 3,300 A. P.

Terra preta Sabemos ahora que la Amazonia no es un bosque primario, sino que fue profundamente modificado durante milenios por el ser humano. A manera de ejemplo, fue éste quien favoreció particulares asociaciones vegetales. Así también desarrolló en especial técnicas agrícolas extremadamente sabias que cambiaron de forma definitiva la imagen del bosque tropical más grande del mundo. Hemos visto anteriormente el caso de los camellones, pero hay otro tipo de agricultura hecha en suelos negros antrópicos llamados terras pretas, cuyas excelente fertilidad se obtiene gracias a la presencia de carbón de madera y restos de actividad humana. La terra preta sería

el resultado de intensivas y largas ocupaciones humanas en un mismo lugar. La estructura porosa del carbón, de elementos orgánicos y de micro vestigios antrópicos permite en efecto retener en el suelo las sales minerales ordinariamente lavadas por la lluvia (Figura 10). Las experimentaciones recientes han demostrado que la

58

Figura 10. Terra preta llena de tiestos de cerámica en el sitio Hatahara, cerca de Manaos, Brasil. © Proyecto Amazonía Central.

fertilidad de estos suelos modificados es de dos a tres veces superior a aquella de las tierras pobres amzónicas. Los conjuntos de terra preta se extienden en general a lo largo de 1 a 5 ha, aunque

algunos sobrepasan las 300. Su espesor promedio es de 40 a 60 cm, pudiendo alcanzar más de 2 m en algunos casos. El conjunto de superficies de terra preta cubriría al menos de 0,1 a 0,3% de la Amazonia (Figura 11).

Los sitios más antiguos de terra preta remontan a 4,800 años, es decir a la época

precerámica, y ciertos investigadores estiman que ésta podría haber comenzado a formarse incluso hace 6,000 u 8,000 años en determinados lugares. Los sitios de terra preta se multiplican y extienden a partir del 500 a. C. y sobre todo durante la segunda mitad del primer milenio d. C. El impacto viral de la conquista europea

provocó una caída demográfica y la disparición de las grandes implantaciones humanas de la cuenca amazónica, deteniendo la formación de terras pretas.

Megalitos Pero los amazónicos no trabajaron únicamente la tierra, puesto que a veces la piedra misma sirvió para trazar signos o delimitar espacios sagrados. Así, pequeñas losas rocosas pueden hallarse alineadas o dispuestas a fin de representar un animal, una lagartija o una tortuga, por ejemplo, como en el caso del “pan de azúcar”, aquellos inselbergs de granito del macizo del Mitaraka, en la frontera sur de la Guayana Francesa con Brasil (Figura 12). En otros lugares a lo largo de la costa de Amapá, inmensas losas de granito de varias toneladas fueron talladas y transportadas a lo largo de kilómetros para luego

59

Sitios monumentales precolombinos en Amazonia ser plantadas en tierra, dibujando una línea, un círculo o un triángulo. Se trata de enormes losas de granito parcialmente talladas, elevadas y dispuestas en línea, círculo o triángulo (Figura 13). Se observa una gran diversidad de formas, dimensiones y disposiciones. Las losas pueden ser plantadas en posición horizontal, vertical o inclinada. Estos sitios estan ubicados en la cima de colinas y tienen un diámetro de 10 a 30 m. La zona delimitada por estos menhires cumplió ciertamente una función

Figura 11. Mapa de la distribución de la terra preta en la cuenca amazónica. © S. Rostain según Clement et al., 2015.

Figura 12. Sitio de piedras alineadas en motivo zoomórfico sobre un inselberg del Mitaraka, frontera entre Guayana Francesa y Brasil. © Servicio Regional de Arqueología de Guayana Francesa.

60

Figura 13. Piedras levantadas formando un círculo en el sitio megalítico Rego Grande, estado de Amapá, Brasil. © S. Rostain. ceremonial importante, como lo sugiere la presencia de sepulturas en urna al pié de las piedras erguidas o el alineamiento preciso de algunas en los solsticios (Petry Cabral & Moura Saldanha, 2009).

Arte rupestre La piedra fue igualmente el soporte de expresiones artísticas variadas, entre las cuales los petroglifos son las más famosas (Figura 14). Estos grabados fueron realizados en rocas a menudo cercanas a los cursos de agua. Si bien es cierto que la escultura con simples herramientas de piedra limita la posibilidad de exuberancia, allí también los amerindios probaron una gran habilidad y mucha imaginación para representar sus diseños. Las figuras se disputan a menudo la superficie de un soporte. Son siluetas humanas, peces, serpientes, aves, monos, caimanes o felinos, que se entrecruzan en medio de diversos signos geométricos en estas arcas de Noé de piedra. Aunque el sentido de estos zoológicos de roca se nos escapa todavía, su frecuencia denota una función precisa y una necesidad recurrente. Cuando los Amerindios no grabaron en las rocas, pintaron rara vez las paredes de los abrigos rocosos. La similitud de los diseños con aquellos de los petroglifos permite pensar en un uso comparable. Este arte empezó con los primeros habitantes de la Amazonia, hace 11 mil años (Pereira, 2003; Jorge et al., 2006).

Emergencia patrimonial Al término de este panorama de terraplenes amazónicos, se destaca una variedad sorprendente de estructuras y modificaciones del paisaje. Hoy los arqueólogos y especialistas de las ciencias de la tierra aceptan estas estructuras como realizaciones precolombinas y no como formaciones naturales, incluso aunque en algunos casos no se puede negar que se trate de estas últimas. La organización y la arquitectura de estas construcciones sugieren un trabajo comunitario, ciertamente bajo la autoridad de un especialista y un poder centralizado. En diversos puntos de las tierras bajas

61

Sitios monumentales precolombinos en Amazonia sudamericanas, la arqueología viene revelando desde hace algunos años indicios de sociedades precolombinas estratificadas. Por lo general su surgimiento aparece alrededor del 500 y el 600 d. C., y habría culminado alrededor del año 1000, cuando comenzó el declive, hacia 1200 d. C., y su desaparición hacia la conquista europea. De cualquier modo es ahora innegable que el hombre ha dejado su huella indeleble en biotopos tan extremos como ciénegas o el bosque tropical. Además de los mecanismos naturales, se deben tomar en cuenta las actividades humanas. Desde la época colonial y hasta la fecha, el fuego, la agricultura y la ganadería, las terrazas y las obras en los camellones elevados asociados a las zonas húmedas aceleran la destrucción de las estructuras de tierra precolombinas. Ésta es la amenaza de origen antrópico que más debemos temer, a menudo acompañada de una destrucción directa a causa de las obras modernas. Los riesgos comprenden intervenciones activas como los drenajes, labrados, obras inmobiliarias o conversiones en campos de arroz. Por ejemplo, por causa de su ubicación en sabanas abiertas donde se hacen varios trabajos modernos (agricultura, ganadería, construcciones, etc.), los campos elevados sufren hoy multiples amenazas (Figura 15). Como es normal, casi todos los países y regiones del mundo conservan monuFigura 14. Roca grabada del sitio mentos humanos de diversa naturaleza inscritos como Patrimonio Mundial. Es Anzú, provincia de Pastaza, Ecuador. menos normal ver todavía a un continente privado de este privilegio. Debido a © S. Rostain.

62

Figura 15. Campos elevados del complejo de Savane Maillard, Guayana Francesa, amenazados por varios peligros: enrasamiento con excavadora a la izquierda, ganado con cerca en el centro, incendio con el área quemada abajo y construcciones con la vía y su fosa lateral a la derecha. © S. Rostain.

su dimensión y diversidad, la Amazonia, el mayor bosque tropical del mundo, es prácticamente un continente, ya que cubre cerca de la mitad de América del Sur (17,840,000 km2). Tiene una superficie comparable con la de Australia. Y es igualmente el resultado de una fuerte interacción de ser humano y naturaleza durante más de 10 mil años. Propusé en este artículo una lista de las obras humanas que, en mi opinión, merecerían estar inscritas en el patrimonio mundial. El análisis de las estructuras de tierra y a veces de piedra precolombinas concluye en una constatación muy alejada de la imagen tradicional de una Amazonia poblada por pequeñas tribus semisedentarias, heredada de una tenaz visión eurocentrista. Si bien la historia precolombina de la Amazonia no fue escrita en los libros, si bien no hubo templos de piedra, los amerindios del bosque tropical inscribieron sin embargo muy claramente sus anales en la tierra. Se vuelve entonces necesario tomar conciencia de este patrimonio original y preservarlo.

Agradecimientos Agradezco a Belém Muriel por la traducción del texto al español; a Eduardo Neves, Martti Pärssinen, Heiko Prümers, Anna Roosevelt y al Servicio Regional de Arqueología de Guayana Francesa, por sus fotografías.

Referencias Balée, W. 1989. The culture of Amazonian forests. En: Posey, D. & Balée, W. (eds.), Resource Management in Amazonia: Indigenous and Folk Strategies, Advances in Economic Botany, 7, The New York Botanical Garden, Bronx, pp. 1-21.

Balée, W. & Erickson, C. L. (eds.) 2006. Time and Complexity in Historical Ecology. Nueva York, Columbia University Press.

Clement, C. R., Denevan, W. M., Heckenberger, M. J., Junqueira, A. B., Neves, E. G., Teixeira, W. G. & Woods, W. I. 2015. The domestication of Amazonia before European Conquest. Proceedings of Royal Society B, Vol. 282, No. 813, doi. org/10.1098/rspb.2015.0813.

63

Sitios monumentales precolombinos en Amazonia Denevan, W. M. 2001. Cultivated Landscapes of Native Amazonia and the Andes. Nueva York, Oxford University Press.

Erickson, C. 2008. Amazonia: the historical ecology of a domesticated landscape. En: Silverman, H. & Isbell, W. (eds.), Handbook of South American Archaeology, Springer/Kluwer/Plenum, pp. 157-183.

Glaser, B. & Woods, W. I. (eds.) 2004. Amazonian Dark Earths: Explorations in Space and Time. Heidelberg, Springer.

Grenand, P. 1982. Ainsi parlaient nos ancêtres: essai d’ethnohistoire Wayapi. París, ORSTOM.

Heckenberger, M. J. 2005, The Ecology of Power. Culture, place and personhood in the southern Amazon, AD 1000-2000. Nueva York, Routlegde.

Iriarte, J., Power, M. J., Rostain, S., Mayle, F., Jones, H., Watling, J., Whitney, B. S. & McKey, D. B. 2012. Fire-free land use in pre-1492 Amazonian savannas. Proceedings of the National Academy of Sciences of the USA, Vol. 109, No. 17, pp. 6473-6478.

Jorge, M., Prous, A. & Ribeiro, L. 2006. Brasil rupestre: arte pré-histórica brasileira. Curitiba, Zencrane Livros.

Lévi-Strauss, Claude, 1994, Saudades do Brasil, Librairie Plon, París. Lombardo, U. & Prümers, H. 2010. Pre-Columbian human occupation patterns in the eastern plains of the Llanos de Moxos, Bolivian Amazonia, Journal of Archaeological Science, 37: 1875-1885.

McEwan, C., Barreto, C. & Neves, E. (eds.) 2001. Unknown Amazon. Londres, The British Museum Press.

Neves, E. G. 2012. Sob os tempos do equinócio: oito mil anos de história na Amazônia

central (6.500 AC – 1.500 DC). Tesis para optar por el grado de Livre-Docente. São Paulo, Museu de Arqueologia e Etnologia/Universidade de São Paulo. Pereira, E. 2003. Arte rupestre na Amazônia: Pará. São Paulo, UNESP. Petry Cabral, M. & Da Moura Saldanha, J. D. 2009. Note sur des structures mégalithiques en Guyane brésilienne. Journal de la Société des Américanistes, Vol. 95, No. 1, pp. 97-110.

Prümers, H. 2014. Sitios prehispánicos con zanjas en Bella Vista, Provincia Iténez, Bolivia. En: Rostain, S. (ed.), Amazonía. Memorias del 3er Encuentro Internacional de

64

Arqueología Amazónica, Quito, 3 EIAA/MCCTH/SENESCYT, pp. 73-89.

Roosevelt, A. C. 1991. Moundbuilders of the Amazon: geophysical archaeology on Marajó Island, Brazil. Nueva York, Academic Press.

Roosevelt, A. C., Housley, R. A., Imazio da Silveira, M., Maranca, S., & Johnson, R. 1991. Eight millenium pottery from a prehistoric shell midden in the brazilian Amazon. Science, No. 254, pp. 1621-1624. Rostain, S. 2011. Los edificadores de la selva: obras precolombinas en Amazonía. En: Chaumeil, J.-P., Espinosa de Rivero, Ó. & Cornejo Chaparro, M. (eds.), Por donde hay soplo, Actes & Mémoires No. 29, Lima, IFEA, pp. 69-87.

——. 2012a. Islands in the rainforest. Landscape management in pre-Columbian Amazonia. Walnut Creek, Left Coast Press.

——. 2012b. Between Sierra and Selva: Pre-Columbian Landscapes in the Upper Ecuadorian Amazonia. En: Catto, N. (ed.), Quaternary International, Special Issue “Human Occupation of Tropical Rainforests”, No. 249, pp. 31-42.

——. 2016. Amazonie. Un jardin sauvage ou une forêt domestiquée. Collection Essai d’écologie historique, Arles, Actes Sud/Errance.

Schaan, D. 2011. Sacred geographies of ancient Amazonia: Historical ecology of social complexity. Walnut Creek, Left Coast Press.

——. 2014. Cronologia das transformações das paisagens amazônicas. En: Rostain, S. (ed.), Amazonía. Memorias del 3er Encuentro Internacional de Arqueología Amazónica, Quito, 3 EIAA/MCCTH/SENESCYT, pp. 51-71.

Steege, H., Pitman, N. C. A., Sabatier, D., Baraloto, C., Salomão, R. P., Guevara, J. E. et al. 2013. Hyper-dominance in the Amazonian tree flora. Science, Vol. 342, No. 6156, pp. 1243092.

Walker, J. H. 2004. Agricultural Change in the Bolivian Amazon. University of Pittsburgh Memoirs in Latin American Archaeology, 13. Pittsburgh/Trinidad, Fundación Kenneth Lee. ——. 2012. Recent Landscape Archaeology in South America. Journal of Archaeological Research, Vol. 20, No. 4, pp. 309-355.

65

Michael Heckenberger University of Florida, USA

Abatract

Resumen

Tropical landscapes in the past were more woven than

Los paisajes tropicales del pasado eran más tejidos que

hammered or cleaved, much like today. They nudged

martillados o escindidos, como hoy. Ellos empujaron

the land through minor manipulations, entwining

la tierra a través de manipulaciones menores, entrela-

the forest into houses and settlements and reworking

zando el bosque en casas y asentamientos y reelabo-

the myriad and inexorable ebbs and flows of water

rando la miríada e inexorable flujo y reflujo de agua

and knitted it all together through human aesthetics

y tricotándolo todo a través de la estética humana y la

and industry. In the Xingu, dense networks of roads

industria. En el Xingu, las densas redes de carreteras

linking towns and villages were hyper-planned and

que unían ciudades y aldeas fueron hiperplanificadas

trees were cycled and recycled on a scale seldom ri-

y los árboles fueron reciclados y reciclados a una esca-

valed in the ancient world. In fact, in terms of net-

la que raramente rivalizaba en el mundo antiguo. De

works, connectivity and traffic patterns, they seem

hecho, en términos de redes, conectividad y patrones

almost sublime in the sophistication of their regional

de tráfico, parecen casi sublimes en la sofisticación de

planning and how it was designed to work with the

su planificación regional y cómo se diseñó para traba-

forest, to nurture rather than tame it, even in compar-

jar con el bosque, nutrirlo en lugar de domesticarlo,

ison to classic urban oases, their historical alter-egos.

incluso en comparación con los oasis urbanos clásicos,

The planning echoes the alternative model of urban

sus alter-egos históricos. La planificación se hace eco

sustainability put forth as a counterpoint to London’s

del modelo alternativo de sostenibilidad urbana pre-

uncontrolled growth by Ebenezer Howard in Garden

sentado como contrapunto al crecimiento incontro-

Cities of To-morrow (1902). The idea seems almost

lado de Londres por Ebenezer Howard en Ciudades

custom-fit for the Amazon, and for many cases from

de Jardines de Mañana (1902). La idea parece casi per-

the Global South, tropical garden cities and, although

sonalizada para el Amazonas, y en muchos casos des-

decimated by Euro-American colonialism, their lega-

de el Sur Global, las ciudades de jardines tropicales y,

cy lives on today in the forests and livelihoods of living

aunque diezmada por el colonialismo euro-americano,

descendants, struggling today in the southern Ama-

su legado vive hoy en los bosques y medios de subsis-

zon’s ‘arc of fire’. Indeed, their alternative variety of

tencia de los descendientes vivos, luchando hoy en día.

pre-modern urbanism, excavated over the course of

El “arco de fuego” del sur de la Amazonia. De hecho,

two decades through participatory work and partner-

su variedad alternativa de urbanismo pre-moderno,

ships, worked with nature not against it and is a living

excavada en el transcurso de dos décadas a través del

example of how to deal with Amazon today in terms

trabajo participativo y las asociaciones, trabajó con la

of biodiversity, sustainability, climate and indigenous

naturaleza no en contra de ella y es un ejemplo vivo

heritage.

de cómo lidiar con la Amazonia hoy en términos de biodiversidad, Y el patrimonio indígena.

66

Xingu Garden Cities: Domesticated Forests of the Southern Amazon These people (the “Pareci nation”) exist in such vast quantity, that it is not possible to count their settlements or villages, (and) many times in one day’s march one passes ten or twelve villages (arranged around a circular plaza), and in each one there are from ten to thirty houses … (and) even their roads they make very straight and wide, and they keep them so clean that one will find not even a fallen leaf.

Antonio Pires de Campos Filho (1862 [1720], author’s translation).

Introduction The meteoric rise of archaeology in Brazil over the past two decades has led to an exponential increase in knowledge about the pre-Columbian in the Amazon. The cherished imagery of primordial nature and society that guided Western scientists for centuries has given way to views that favour dynamic change in coupled human-natural systems. The deep antiquity and diversity of occupations across the world’s largest tropical forest biome, including large, settled and regionally integrated complex societies, rivals other large forest biomes across the globe. The view of the area as a vast natural laboratory, however, lingers on in scientific discussions about the Amazon, as if archaeologists are simply misguided, at least in their reach, and the extrapolation of settled, agricultural life too broadly. Thus, the pendulum swung too far to the left, to the humanistic and historical, from the hard-nosed business of natural science. The deep antiquity and diversity of occupations across the world’s largest tropical forest biome, including large, settled and regionally integrated complex societies, rivals other large forest biomes across the globe.

67

Xingu Garden Cities: Domesticated Forest of the Southern Amazon As our understanding of its rich cultural heritage increases, the Amazon region itself is undergoing a radical transformation due to rapid development, as forests are ground into the dirt, inundated and paved-over. Needless to say, this acutely threatens biodiversity, most particularly the integrity of ecological systems that have built up over millennia. As we learn more about what is hidden under the forest canopy, often visible as the result of the very development that also threatens to destroy these cultural heritage resources, it is clear that, far from timeless, the Amazon was no less dynamic, and its people no less agential or ingenious than anywhere else. Notably, the Late Holocene saw the development of large, densely settled regional societies, often as hierarchically organized peer polities across the region. The discovery of Amazonian forms of complex and intensive land-use has far-reaching indirect impacts on biodiversity and human population densities, including catastrophic depopulation in early colonial times across the region (Clement et al., 2015). This view has its critics, but virtually all anthropologists agree that there were some remarkably complex systems, at least those working directly with archaeological evidence, historical archives and ethnographic and linguistic evidence of indigenous history. Even areas where impact is slight or indirect, humans changed the face of the Amazon through subtle modifications throughout the Late Holocene. This suggests very different pathways of socio-political complexity and economic intensification, land-use and relations between humans and nature than might be expected by Western scholarship and world historical schema. Indigenous knowledge among Amazonian peoples, as a group, is quite distinctive to Western scientific knowledge; indeed, it is the polar opposite or alter ego of it according to many regional specialists since Lévi-Strauss (1961). These past complex societies shared many things with the traditional Amazonian today: people ate insects as often as large herbivores, palm fruits and tubers rather than cereal crops, used bows and blowguns, went around naked, painted and feathered, among a host of other things. Many are still in use or directly remembered by the descendant indigenous groups today. They favoured diverse systems of plant and animal management, with little more than a digging stick and occasional burning, focusing on tree crops, including orchards and agro-forestry, vegetative root crops and house gardening in rich terra preta compost, rather than ploughs, draft animals, ‘sowing’ of seed or other intensive systems of mono-cropping. Indeed, the over 80 native species in some state of domestication (Clement et al. 2010) and almost dizzying array of different varieties recognized by indigenous peoples suggest that agro-biodiversity is as varied as biodiversity. The majority of managed plants are trees, numerous fruits, including palms, which are also singular as industrial crops in agro-forestry systems. In this world of perishable technologies, it seems everything is woven from forest products, save the rare ceramics, even rarer stone and metals, from the smallest baskets, hammocks, fences, weirs, houses and even whole villages, the ring of houses being like the selvage of a basket. Certainly, weaving is a better metaphor for the overall architecture, the built environment, rather than the building. In today’s pragmatic world, these organic built environments seem far more ‘user-friendly’ or as the cliché goes, ‘at

68

one with nature,’ than what passes for techno-economic achievements elsewhere, hard surfaces and farmland. Like the Global South and the tropical world, more generally, these forest civilizations need to be considered on their own terms, since their historical development is not easily measured according to a yardstick of pre-existing Western historical experience. This paper considers one feature still commonly denied: preColumbian urbanism and the attendant complex land-use and production systems associated with urbanization. What might Amazonian urbanism look like? What are we looking for? Remarkably, one of the most compelling models of what Amazonian urbanism might have been like was presented by Ebenezer Howard’s utopian vision of a sustainable urban development, Garden Cities of To-morrow

(1902), in which he proposed to curb the degradation of cities, such as London, where he lived, on the eve of the twentieth century, the peak of the Industrial Era. Howard was the forefather of what today would be called ‘urban sustainability’ – at the vanguard of the first generation of the ‘green movement.’ He proposed an alternative model of urban development by design, specifically an urban fabric, constructed from scratch, which was not based on singular cities or large centres, but instead in networks of towns, integrated by roads that extend across mosaic green spaces, overall promoting a mixed and patchy urban landscape. Had Howard known of the Upper Xingu in the southern Brazilian Amazon, he might have added a passage to his book, ‘garden cities of yesterday.’ These Amazonian garden cities, or galactic urbanism, are comprised of a multi-centric network of ancient walled-towns, satellite villages and hamlets, heavy and wellcontrolled road and waterway traffic, within an overall system of land use and spatial planning. It documents an almost sublime complexity in design, seeming to link everything to everything else, which ultimately worked with the forest, rather than simply cutting it down, by subtly reworking the contours of the land and altering the proportion of useful species and areas.

Domesticated Forests In The ‘Arc Of Fire’ Today, archaeologists agree that twentieth century models of limited human potential and sparse population fail to capture the substantial cultural and historical variation now known to exist across all parts of Amazonia. Over the past two decades, a consensus has emerged among regional specialists that many parts of Amazonia were anthropogenic, substantially modified by long-term human occupations throughout the Holocene and, particularly, the influence of large, settled and regionally organized late pre-Columbian populations in some areas (Heckenberger and Neves, 2009). A precipitous increase in archaeological research since the mid-1990s reveals diverse complex societies and dynamic human-natural systems, including dozens of cases of large pre-Columbian and early historic settlements (greater than 50 ha), high regional site densities and substantial human intervention and even engineering of tropical forest natural environments.

69

Xingu Garden Cities: Domesticated Forest of the Southern Amazon Recently, palaeobotanists have rightly cautioned against extrapolating the anthropogenic model too broadly, based primarily on a small sample of soil augers from interfluvial areas of western Amazonia (for example, Bush et al., 2015; McMichael et al., 2012; Piperno et al., 2015). This does not imply that all areas were so influenced or even that most of the Amazon forest was strongly influenced by humans. It does recognize enclaves of densely settled pre-Columbian populations have been identified from all major subregions, which constructed vast ‘domesticated’ landscapes through their semi-intensive land management practices. It also suggests that such pronounced human impacts had far-flung indirect and subtle effects on nature, including the fact that cultural choice, socio-political relations and historical factors were potentially as determinant of where relatively untouched forest existed. This debate clearly has broad implications for biodiversity, ecological resiliency and climate change in the world’s largest forest. However, the polarized and intransigent positions are hard to reconcile due to the lack of comparable datasets and problem-oriented interdisciplinary research, including indepth field archaeology and palaeobotanical analysis of well-controlled samples. Underlying critiques of historical ecology and an anthropogenic Amazon follow the assumption that forests will be converted to farmland in cases of larger, settled populations, a pattern that can be captured by minimal sampling. In Amazonia, archaeology and plant studies suggest that systems of human-landplant management are highly diversified in terms of land-use. Rather than the typical Neolithic model of the rise of fixed plot agricultural practices typical of the Neolithic and ‘intensification’ through the use ploughs and draft animals, irrigation and mono-cropping these systems rely on a vast inventory of managed plants, with over 80 species in some stage of ‘domestication’ and a dizzying array of named varieties in indigenous knowledge systems (Clement et al, 2015; Neves, 2016). Specifically, studies are needed to determine the parameters of different systems of forest conversion (or not), in terms of highly variable human impacts on soils and vegetation within diverse forest management systems, notably including the ‘intermediate disturbance’ implied in many archaeological cases (Balée, 2013). Complex built environments are widely known in riverine areas of the southern headwaters. In the eastern Llanos de Mojos and along the Guaporé River, complex system of earthworks, including causeways, fish weirs and ponds and forest islands (ancient settlements), raised fields and diverse other archaeological features are documented in extensive ‘domesticated landscapes’ of forested areas (for example, Denevan, 2001; Erickson, 2006; Walker, 2004). Recent work in south-western Amazon has shown dynamic change, including substantial human interventions in the form of ‘geoglyphs,’ ditches and low linear mounds creating enclosures and causeways, which likely connect these over 250 sites across the region (Schaan, 2011), tied to forest conditions in these transitional forests that fluctuated in some cases between the broadleaf evergreen Amazonian forests and the more open plains of central Brazilian cerrado (Carson et al., 2014, 2015).

70

Figure 1: Vegetation map showing primary forest types in Brazil (from Heckenberger et al. 2008). The numbered white circles represent areas of well-documented dense, regionally organized populations, including in the southern Amazon transitional forests the Upper Xingu (1), Upper Tapajós (2) and middle Guaporé (3) areas. © Michael Heckenberger

The southern Amazon transitional forests are an important macro-regional case, since there is a very high degree of alteration in areas of major pre-Columbian occupations, which extend across much of the region. Pre-Columbian regional systems conformed to the broad contours of the natural environment, but major enclaves were concentrated in areas of dense standing tropical forest, creating ‘saturated anthropogenic landscapes’ and domesticated forests (Figure 1). Like the better known floodplain polities along the Amazon, human interventions transformed the forests of across the southern Amazon transitional forests, a macroecological province of tropical forest between the Amazonian broad leaf evergreen forests and xerophytic open woodlands and savanna (cerrado) of central Brazil. These societies have a common origin in the Arawak-speaking groups spread across the lowlands, which settled into the major river valleys and dominated these large forested areas, in nearly one-to-one correspondence. This underscores the coupling of cultural and ecological variability and the need to address human and natural factors to understand the composition and functioning of biodiversity, particularly considering the degree to which eco-systemic biodiversity was linked to human interventions. This in turn influenced plant and animal species and genetic variation and their dynamic changes over the past millennium. In the southern peripheral tropical forests of the Brazilian Amazon, the two headwater basins of the Xingu and Tapajós River were home to large, settled pre-Columbian peer polities across regions co-extensive with the transitional forests of the basins,

71

Xingu Garden Cities: Domesticated Forest of the Southern Amazon marking a pronounced ecological boundary between the high Amazonian forests and the open landscapes of Central Brazil. It also suggests complex socio-historical diversity in all areas, including areas of small size and low density of occupation sites in sparsely populated hinterlands, organized in concentric circles of human influence in regional systems, with ‘no-man’s lands, buffer zones and ‘preserves.’ Closed forest conditions prevail today in areas that were dominated by large, settled and regionally organized polities in pre-Columbian times, and continue ot be occupied by descendant indigenous groups, making the question not what is anthropogenic in these forests but what isn’t. The real job ahead lies in creating well-controlled case studies of what exactly the Amazon was like in pre-Industrial times. There is an urgent for far more detailed analyses on any part of the Amazon, in order to determine forest cover change and resource management, notably during late pre-Columbian times. What types of land management, including forest conversion, were associated with different socio-economic regimes, such as extensive slash-and-burn or intensive fixed plot agriculture or, as proposed here, complex hybrid systems, which were more focused on forest management than crop production focused on a few staple crops in fixed rotational cycles? The Upper Xingu, presented here as an exemplary case of southern Amazonian domesticated landscapes, allows the suggestion that the entire forested area within clusters were like big gardens, partitioned off into agricultural zones with their own internal cycles, including many areas that were occasionally denuded small patches of forest in long-term, inter-generational cycles which generally promoted forest cover.

Upper Xingu The Upper Xingu is the easternmost of the six or eight of these large regional cultures in the southern transitional forests. Archaeological research over the past twenty years has documented discrete late pre-Columbian and early historical occupations, as well as a continuous record of habitation by related (Xinguano) peoples over the past millennium (Note 1). The Xingu entered written history only in 1884, when indigenous populations were already greatly reduced (some 3,000-4,000 people in over 30 autonomous villages). By the time the indigenous reserve, the Parque Indígena do Xingu (PIX), was established in 1961, the Xinguanos were reduced to a little over 500 persons, in a dozen relatively autonomous and isolated settlements. These were the remnants of the oncethriving late pre-Columbian regional peer polity, which was spread across nearly 100 km², likely numbering well over 50,000 inhabitants. It was composed of several dozen discrete multi-settlements clusters, organized into tightly integrated polities. Campos’ (1862) description of the closely related ‘Pareci nation’ in the neighbouring Upper Tapajós River in 1720, quoted above, seems tailor-made as a benchmark for the eclipse of the regionally integrated peer polities, which were already drastically reduced from the heyday of the late pre-Columbian polities, largely due to epidemics.

72

Archaeological research in the Xingu region provides some clear evidence of large, densely settled late pre-Columbian populations in the lowland tropical forests (for example, Heckenberger, 2005, 2010, 2013; Heckenberger et al., 2003, 2008, 2013; Note 2). Archaeological studies, alongside oral history and historical ethnography and linguistic reconstructions document several major periods of transformational change in the evolving regional system (supported by 30 C14 dates): (1) Initial colonization by settled agriculturists, c. 1500 BP or before, ancestral to later Xinguano Tradition); (2) a ‘Galactic Period,’ with enlargement and

structural elaboration of major settlements, c. 1350 to 1650, marked by large-scale constructions (roads, canals, weirs, ponds, defensive walls and other engineered features) and inferred integration in hierarchical clusters; (3) population decline and fallowing of the forest landscapes, after c. 1650 (Proto-Historic Period); (4)

geographic compression from depopulation and ethnogenesis of multi-lingual contemporary society (early Xinguano Period), after c. 1750, including population stabilization based on scale-adjusted settlement pattern or one or few small villages

related to individual named groups (otomo) in regional socio-cultural system; and (5) a period of acute depopulation after Xinguano peoples enter the written record (1884) followed by post-1980 demographic and socio-economic rebound and restabilization and current growth of all ten autonomous descendant groups, and increasing political autonomy (late Xinguano Period). Indigenous research assistants were critical to identify historical villages (etepe) and ancient settlements (ingiholó-ìtupe), locating earthworks (gepugu), anthropogenic dark earth (egepe) and related distinctive vegetation and ceramics (egeho) that are all closely correlated (Heckenberger, 1996, 2009; Schmidt, 2010). Once located, all sites were positioned with real-time GPS or on satellite images (in rare cases where GPS was unavailable). Basic mapping of major earthworks was conducted in and around 14 sites (n=14) and detailed mapping of earthworks was conducted at eight sites (X6, X13, X11, X17-20, X22). Earthworks included peripheral ditches with berms and low linear berms (.5 to 1.5 m high) situated at the margins of major roads (greater than 10 m wide) and circular public plaza areas. These constitute roughly 40 km of continuous archaeological features, including curbed roads between settlements, as well as gate, bridge and weir areas at site margins. Mapping of major earthworks at these sites reveals an elaborate regional plan, including major excavated ditches surrounding the largest settlements (up to 15 m wide, 5 m deep, and 2.5 km in length), linear mounds positioned along roads and public plaza areas, and a variety of wetland constructions, such as raised causeways, bridges, river obstructions (weirs), canals and artificially modified ponds. Two integrated clusters of walled towns and non-walled villages were identified. These are inferred to be small, independent polities occupying territories of over 250 km² in an integrated regional peer polity of at least 25-30 such clusters. In the better-known northern (Ipatse) cluster, settlements were positioned at forest/ wetland interfaces at regular intervals (3-8 km) and linked by a region-wide system of broad, straight roads. Clusters were composed of large- (≥40 ha) and

73

Xingu Garden Cities: Domesticated Forest of the Southern Amazon medium-sized (

Smile Life

When life gives you a hundred reasons to cry, show life that you have a thousand reasons to smile

Get in touch

© Copyright 2015 - 2024 PDFFOX.COM - All rights reserved.