You are on page 1of 53

Tropical Forests of the Guiana Shield

Ancient Forests in a Modern World


This page intentionally left blank
Tropical Forests of the Guiana Shield

Ancient Forests in a Modern World

Edited by

D.S. Hammond

Formerly of Iwokrama International Centre for Rain Forest


Conservation and Development
Georgetown, Guyana

CABI Publishing
CABI Publishing is a division of CAB International

CABI Publishing CABI Publishing


CAB International 875 Massachusetts Avenue
Wallingford 7th Floor
Oxfordshire OX10 8DE Cambridge, MA 02139
UK USA

Tel: +44 (0)1491 832111 Tel: +1 617 395 4056


Fax: +44 (0)1491 833508 Fax: +1 617 354 6875
E-mail: cabi@cabi.org E-mail: cabi-nao@cabi.org
Web site: www.cabi-publishing.org

©CAB International 2005. All rights reserved. No part of this publication may be
reproduced in any form or by any means, electronically, mechanically, by
photocopying, recording or otherwise, without the prior permission of the
copyright owners.

A catalogue record for this book is available from the British Library, London,
UK.

Library of Congress Cataloging-in-Publication Data


Tropical forests of the Guiana Shield: ancient forests of the modern world / edited
by D. Hammond
p. cm.
Includes bibliographical references and index.
ISBN 0-85199-536-5 (alk. paper)
1. Rain forest ecology--Guyana Shield. 2. Rain forests--Guyana Shield. 3. Natural
History--Guyana Shield. 4. Guyana Shield. I. Hammond, D. (David S.) II. Title.

QH111.T76 2005
578.734´098--dc22
2004021133

ISBN 0 85199 536 5

Typeset by MRM Graphics Ltd, Winslow, Bucks


Printed and bound in the UK by Biddles Ltd, King’s Lynn
Contents

Contributors vii

Acknowledgements ix

Acronyms and Abbreviations xi

1 Ancient Land in a Modern World 1


David S. Hammond

2 Biophysical Features of the Guiana Shield 15


David S. Hammond

3 Ecophysiological Patterns in Guianan Forest Plants 195


Thijs L. Pons, Eustace E. Alexander, Nico C. Houter, Simmoné A. Rose
and Toon Rijkers

4 Rainforest Vertebrates and Food Plant Diversity in the Guiana Shield 233
Pierre-Michel Forget and David S. Hammond

5 Folivorous Insects in the Rainforests of the Guianas 295


Yves Basset, Neil D. Springate and Elroy Charles

6 Flower-visiting Insects in Guianan Forests: Pollinators, Thieves, Lovers and


Their Foes 321
Bart P.E. De Dijn

7 Guianan Forest Dynamics: Geomorphographic Control and Tropical Forest


Change Across Diverging Landscapes 343
David S. Hammond

8 Socio-economic Aspects of Guiana Shield Forest Use 381


David S. Hammond

v
vi Contents

9 Forest Conservation and Management in the Guiana Shield 481


David S. Hammond
Index 521
Contributors

Alexander, E.E., Planning and Research Development Division, Guyana Forestry


Commission, 1 Water Street, Kingston, Georgetown, Guyana.
Basset, Y., Smithsonian Tropical Research Institute, Apartado 2072, Balboa, Ancon,
Panama. (e-mail: bassety@tivoli.si.edu)
Charles, E., Faculty of Agriculture/Forestry, University of Guyana, Turkeyen, Georgetown,
Guyana. (e-mail: elroy_c@yahoo.co.uk)
De Dijn, B.P.E., National Zoological Collection, University of Suriname, University
Complex, Leysweg, PO Box 9212, Paramaribo, Suriname. (e-mail: nzcs@cq-link.sr)
Forget, P.-M., Département Ecologie et Gestion de la Biodiversité, Museum National
d’Histoire Naturelle, UMR 5176, CNRS-MNHN, 4 Av. Du Petit Chateau, F.91800
Brunoy, France. (e-mail: forget@mnhn.fr)
Hammond, D.S., Iwokrama International Centre for Rain Forest Conservation and
Development, Georgetown, Guyana. Current Address: NWFS Consulting, 15595 NW
Oak Hill Dr., Beaverton, OR 97006, USA. (e-mail: dhammond@nwfs.biz)
Houter, N.C., Department of Plant Ecophysiology, Utrecht University, Sorbonnelaan 16,
3584 CA Utrecht, The Netherlands.
Pons, T.L., Department of Plant Ecophysiology, Utrecht University, Sorbonnelaan 16, 3584
CA Utrecht, The Netherlands. (e-mail: t.l.pons@bio.uu.nl)
Rijkers, T., Department of Forest Ecology and Forest Management, Wageningen University,
PO Box 47, 6700 AA Wageningen, The Netherlands.
Rose, S.A., Planning and Research Development Division, Guyana Forestry Commission, 1
Water Street, Kingston, Georgetown, Guyana.
Springate, N.D., Department of Entomology, The Natural History Museum, Cromwell
Road, London SW7 5BD, UK. (e-mail: nds@nhm.ac.uk)

vii
This page intentionally left blank
Acknowledgements

Any work is the product of numerous contributions at many different moments in many
different ways. This book is no different and the individuals who have contributed through
their patient and persistent collection of data in the field, organization and compilation of
references, time series and spatial information and administrative support are numerous.
I would particularly like to thank again Kate Lance, Kwasie Crandon, Roxroy Bollers,
George Roberts, Catherine Clarebrough, Luc van Tienen, Mariska Weijerman, Arnoud
Schouten and Dexter Angoy for their efforts in collecting data presented or referred to in
various chapters of the book. Several UK and Guyanese volunteer and training groups did
a marvellous job with data collection. Twydale Martinborough deserves special thanks for
her efforts in meticulously compiling much of the data on production statistics and
Chanchal Prashad for her excellent work in digitizing and compiling many of the geologi-
cal and historical GIS coverages presented in figures throughout the book. Thanks are
extended to all of the scientists and resource management professionals in French Guiana,
Guyana and Suriname for sharing information and literature early in the book’s develop-
ment. I also owe a debt of gratitude for the efficient and timely administrative support of
Juliet Dos Santos and Jean Bacchus in making the compilation of ‘grey’ literature and sta-
tistics that much easier. The amazing efficiency and speed through which high quality data
were made available by governmental institutions in Guyana, Venezuela, Suriname,
France, Brazil, the USA and the UK, often free-of-charge and through ftp downloading, has
enhanced in no small measure the quality of content and avoided the labyrinth typically
encountered en route to acquiring such types of environmental data. I would like to
acknowledge their effort here as a means of registering support for the growing trend in
making these types of data available to the international community. These include
ANEEL (Brazil), CDC (USA), CNRS (France), CPRM (Brazil), DAAC (USA), DANE
(Colombia), GEBCO (UK), IBAMA (Brazil), IBGE (Brazil), INMET (Brazil), IRD (France),
INSEE (France), MARNR (Venezuela), NOAA (USA), NODC (USA), OCEI (Venezuela),
ODP (USA), SCOPE (USA), UEA-CRU (UK), USGS (USA), and Woods Hole Oceanographic
Institute – LBA (USA). Similarly, I wish to thank agencies and commissions under the UN
umbrella – the FAO, UNEP, UNESCO, UNDP, IPCC and WCD, for making data and litera-
ture they have generated widely and easily available.
Administrative, logistical, data-sharing and library support were provided to the book
at various stages by the Tropenbos-Guyana Programme, Tropenbos Foundation, Guyana
Forestry Commission, Geology and Mines Commission, Imperial College – Silwood Park,
UK Natural History Museum, CABI Bioscience, and the Iwokrama International Centre for
Rain Forest Conservation and Development. In particular, thank you for the great cooper-
ation and exchange with various institutions in Guyana, including the Guyana Forestry

ix
x Acknowledgements

Commission, University of Guyana, National Agricultural Research Institute and


Environmental Protection Agency.
Without the financial support provided to the editor and for project development, this
book would not have been possible. To this end, generous support provided at various
early stages by the Department for International Development-UK, the Tropenbos
Foundation and the European Commission (to the Iwokrama Centre) is gratefully acknowl-
edged. Support during the final stages was provided solely by the editor.
I wish to acknowledge the professional and motivational support provided at various
times through the development of this book by Professor Val Brown (now at University of
Reading), David Cassells (Iwokrama/World Bank), Ben ter Welle (Tropenbos/Utrecht
University/GTZ), my friend and colleague Pierre-Michel Forget (MNHN – Paris) and Hans
ter Steege (National Herbarium – Utrecht). I also would like to thank the many scientific
colleagues who agreed to review chapters at various stages and often at short notice and
the contributing authors for their unprecedented patience while this volume went through
various changes.
The editor would like to dedicate his effort on this book to the memory of Timothy C.
Whitmore, one of the great tropical botanists, ecologists and foresters of the 20th century
and a scientific guiding light during my time living and working in Guyana.
Most of all, thanks to my family for their endless patience.

About the Editor

David S. Hammond has been researching and working to help conserve and sustainably
develop neotropical forests since 1987. He received his BSc in Botany–Environmental
Science from Miami University, USA, and a PhD in Environmental Sciences from the
University of East Anglia, UK. He currently resides in Portland, Oregon.
Acronyms and Abbreviations

ANEEL Agencia Nacional de Energia Eletrica


AVHRR Advanced very high resolution radiometry
BP Before present (normally 1950)
CAM Crassulacean acid metabolism
COADS Coupled ocean–atmospheric data set, NOAA
CVG Corporacion Venezolana de Guayana
DAAC Distributed Active Archive Center, NASA
DANE Departamento Administrativo Nacional de Estadistica, Colombia
DNPM Departamento Nacional de Produção Mineral, Brazil
DOC Dissolved oxygen concentration
DWIC Dutch West Indies Company
ENSO El Niño – southern oscillation
ETR Electron transport rate
FAO Food and Agriculture Organization of the United Nations
GDP Gross domestic product
GFC Guyana Forestry Commission
GGMC Guyana Geological and Mines Commission
GOES Geostationary operational environmental satellite
HIV Human immuno-deficiency virus
Hydromet Guyana Hydrometeorological Office
IBAMA Instituto Brasileiro do Meio Ambiente e dos Recursos Naturais
Renováveis
IBGE Instituto Brasileiro de Geografia y Estatistica
INDERENA Instituto de Desarrollo de los Recursos Naturales, Colombia
INPARQUES Instituto Nacional de Parques, Venezuela
INRA Institut National de Recherche Agronomique
ISR Incoming short-wave radiation
ITCZ Inter-tropical convergence zone
JERS Japanese earth resources satellite
LAPD Latin American Pollen database
LAR Leaf area ratio
LGM Last Glacial Maximum
LMA Leaf mass per unit area
LMF Leaf mass fraction
NAR Net assimilation rate

xi
xii Acronyms and Abbreviations

NCAR National Center for Atmospheric Research


NCEP National Center for Environmental Prediction, NOAA
NEP Net ecosystem productivity
NOAA National Oceanic and Atmospheric Administration
NPV Net present value
NTFP Non-timber forest product
OCEI Oficina Central de Estadistica e Informatica, Venezuela
OLR Outgoing long-wave radiation
ONF Office National de Forêt, France
ORSTOM Institut de Recherche Scientifique pour le Développement en
Coopération
PFD Photon flux density
PS II Photosystem II
RGR Relative growth rate
RIL Reduced impact logging
SAR Synthetic aperture radar
SLP Sea-level atmospheric pressure
SOI Southern oscillation index
SPC Spare productive capacity
SST Sea-surface temperature
TATE Trans-Amazonian tectonothermal episode
TRMM Tropical rainfall measuring mission
TSS Total suspended solids
TZ+ Total dissolved cation concentration
USGS United States Geological Survey
VPD Vapour pressure deficit
WBR World base reference
WCMC World Conservation Monitoring Centre
WMSSC World Monthly Surface Station Climatology, NCAR
WPA World petroleum assessment
WRI World Resources Institute
WUE Water use efficiency
9 Forest Conservation and Management in
the Guiana Shield

David S. Hammond
Iwokrama International Centre for Rain Forest Conservation and Development,
Georgetown, Guyana. Currently: NWFS Consulting, Beaverton, Oregon, USA

Exposed Precambrian Landscapes – a typified by Precambrian shield interiors


Pretext for Caution (see Chapters 7 and 8). Hence, it is not
unreasonable to state that existing archaeo-
More than 550 million years of weathering, logical evidence points to a ‘ring of early
with few or no major diastrophic events, lowland civilization’ encircling the Guiana
can take its toll on the roots of landscape Shield along the major sedimentary basins
carrying capacities. Most available informa- separating it from the Andean highlands
tion for the Guiana Shield converges along and other shield regions south of the equa-
a central theme – ancient surface geology tor (see ‘Human Prehistory of the Guiana
creates difficult conditions for human soci- Shield’, Chapter 8). Ceramic evidence is
eties. These difficult conditions are not cre- widely distributed across the region, but
ated as a result of catastrophic impacts most submitted dates for those located in
commonly associated with risk and calami- the interior are of more recent, even Indo-
tous loss, such as volcanic eruptions, earth- Hispanic ages (see ‘Material types: ceram-
quakes, hurricanes and landslides. Rather, ics’, Chapter 8). Together, this reflects on
it is their absence that created conditions the existing debate over levels and limits on
challenging early human society and its pre-Columbian, lowland cultural develop-
advance towards more sedentary and per- ment and sophistication in one important
manent lifestyles. way. It suggests that the upland terra firme
The nexus of debate concerning the forest regions may have exerted a relatively
prehistory of human settlement across the pronounced limiting effect on human social
Amazon in many ways reflects the transi- structures, but that this was likely conse-
tion across geomorphographic regions. quent, not parallel, to the development of
Accrued artefactual evidence supporting sophisticated, sedentary social centres in
complex social units and long-term inhabi- the main sedimentary depressions across
tation in the lowland neotropics is drawn the neotropics. Evidence also indicates that
almost exclusively from ramping system longer-term occupation of the shield inte-
regions (see ‘Ramping and dampening sys- rior, and its more difficult environmental
tems’, Chapter 7), while depictions of pre- conditions, occurred as groups fled inter-
Columbian lowland forest inhabitants as tribal or inter-clan conflict and later, the
small bands of para-nomadic wanderers sweeping spread of colonizing Europe.
appears to have been strongly shaped by Thus, historical anthropological characteri-
evidence derived from dampening systems zations post-AD 1500 probably reflected

© CAB International 2005. Tropical Forests of the Guiana Shield (ed. D.S. Hammond) 481
482 D.S. Hammond

upon only a very small remnant of the cul- Chapter 2). In part, this is due to a much
tural depth and complexity of lowland soci- higher fraction of area affected by high-
eties that existed along the shield perimeter energy fluvial systems, hurricanes, volcanic
more than 600 years ago. ejecta and landslips across ramping system
Globally, most known centres of early regions, such as Central America and west-
sedentary human civilization are also ern Amazon (Räsänen et al., 1987; Kalliola
noticeably absent from the Precambrian et al., 1999) compared to shield regions.
shield regions of the world, having estab- However, highly infertile and relatively fer-
lished instead in the more suitable tile soil facies can be found across all
Phanerozoic sedimentary basins or moun- regions. The difference is met when consid-
tainous landscapes. In the adjacent ering the relative distribution of facies
Brazilian Shield region, putative pre-Clovis across the fertility spectrum. In this case,
artefacts have been discovered principally soils of the Guiana Shield are highly right-
along the scattered Palaeozoic sedimentary skewed towards uniformly low nutrient sta-
cliffs delimiting the São Francisco water- tus, relative to other regions (see Fig. 2.15).
shed, not in the heart of the adjoining Even floodplain soils offer relatively fewer
shield region (see Fig. 8.1). During the later incremental gains in soil fertility, in contrast
Neo-Indian period, signs of advanced to the well-established differences between
Amerindian civilization are best known varzea soils and adjacent areas of terra firme
from the raised field complexes of sedi- in the downwarp and sub-Andean sedimen-
ment-filled wetlands, the Llanos de Mojos tary basins, as well as across parts of Central
of Bolivia (Denevan, 1970), Llanos de America (see Fig. 2.12).
Barinas of Venezuela (Spencer et al., 1998)
and the sedimentary coasts along the
Atlantic (see ‘Site types – earth engineer- Acidity
ing’, Chapter 8). But was such a vast area
immune from advanced human coloniza- From a hydrochemical perspective, the
tion because of a simple contrast between Amazon basin is a very large, spatially
soil productivity conditions attached to anisotropic, acid–base titration. Acidity is
terra firme and varzea forests? Probably not. regulated westward by the relatively cold,
Tropical forests in Precambrian landscapes alkaline base-rich contribution of the
may have been less susceptible to early Andean slopes, but on average becomes
human colonization for a number of rea- increasingly acidic eastward across the
sons, of which soil fertility counts impor- lowlands, particularly in the Precambrian
tantly. However, not all factors limiting shield regions (see Chapter 2). A large part
human colonization across shield regions of this change-over comes with the duration
are immediately attached to a general, but of exposure to much higher temperatures in
hopefully dwindling (Moran, 1995), mis- the lowlands. The interesting prospect of
conception of uniform upland soil infertil- lower ambient temperatures across the Last
ity across the lowland tropics (Meggers, Glacial Maximum (Liu and Colinvaux,
1996). 1985; Colinvaux et al., 1996) (also see
‘Prehistoric climates of the Guiana Shield’,
Chapter 2) would suggest that these acidity-
Soil fertility buffering contributions would be sustained
in much higher concentrations further
Soil fertility throughout the wet tropics is, downstream during colder phases, creating
however, relatively low in comparison to with it expanded zones of more modest
drier or colder regions of the planet due to acidity and higher soil fertility across west-
much higher acidity and chemical weather- ern Amazon and the downwarp.
ing rates, but relevant differences in soil Anticipating this type of extension across
nutrient content exist across geomorpho- most of the eastern shield regions would
graphic regions (see ‘Soils and soil fertility’, appear much less plausible.
Forest Conservation and Management 483

The southward location of the periphery. In the interior, they are virtually
Brazilian Shield creates a drier seasonal absent, being replaced by some of the most
environment that modulates to some extent acidic fluvial systems known from the trop-
the impacts of high rainfall on acidity ics. Combined with high humic acid con-
under closed-forest conditions. tents in many rivers, acidity and low
Development of the Amazon Downwarp oxygenation causes aquatic productivity to
surfaces from Andean-derived materials plummet and with it the carrying capacities
also assists in buffering these areas against it may have offered to sedentary, prehis-
high acidity that is both spatially wide- toric societies (see ‘River, lake and tidal sys-
spread and persistently low across the tems’, Chapter 2). Only near the turbulence
Guiana Shield. Similarly, across many created by rock features cross-cutting river
regions of Central America and the channels is productivity believed to
Caribbean, more seasonal climate combined increase significantly. These ‘falls’ may
with contemporary weathering of wide- have provided the greatest opportunities for
ranging clastic, chemical (e.g. oolitic lime- long-term occupation in an environment
stones) and organic-based (e.g. fossiliferous otherwise disfavouring sedentarism.
limestones) sedimentary features work to
counter widespread tropical acidity.
Prospects for long-term, prehistoric Access
occupation at the centre of neotropical ter-
restrial acidity, the interior Guiana Shield, Perhaps of equal or greater importance is
would seem counterintuitive when more the likelihood that forest-covered
amenable environments exist along the Precambrian areas posed more substantial
periphery. Acidity is not only linked to ter- challenges to early settlers due to difficul-
restrial productivity through regulation of ties in access. Across the lowland neotrop-
available phosphorus and the concentration ics, people have traditionally traded,
of heavy-metal concentrations that can cre- migrated and communicated along rivers.
ate further physiological challenges to plant Yet, rivers draining shield regions become
survival and growth (see Chapter 3). It also some of the most difficult to navigate as
regulates aquatic productivity and struc- sedimentary cover thins and exposed
tures the trophic assemblage of primary Proterozoic structures create tremendous
producers, consumers and predators. barriers to movement. The role of the ‘falls’
Aquatic productivity is arguably the has been one of historically curtailing more
most important determinant of long-term rapid use of resources in the Guiana Shield
human occupation in tropical environ- interior and concentrating development
ments. As the main source of protein, fish along the margins (e.g. McTurk, 1882; Im
play a pivotal role in sustaining subsistence Thurn, 1883; Goulding et al., 1988).
and buffering more tenuous productivities
attached to upland tropical soils. Across the
varzea regions of the Amazon and its major Control
tributaries, high aquatic productivity is the
one attribute that distinguishes these areas Yet, today those same barriers are increas-
from others. Perhaps only the presence of ingly proving surmountable. Road con-
engineered earth in the form of terra preta struction and an expanding network of
strikes equivalence with relative fertility of aerodromes is increasingly connecting the
the varzea across the eastern Amazon. Yet, shield interior and perimeter and with this,
these too appear largely restricted to the both the prospects and problems of modern
Phanerozoic sediment cover of the Amazon forestland use. Improved and expanded
Downwarp and Atlantic rim of the Guiana access to the interior forested regions of the
Shield see (Fig. 8.2). Guiana Shield has brought forest fire igni-
Across the Guiana Shield, varzea-like tion sources (e.g. Hammond and ter Steege,
features are also largely restricted to the 1998), an upsurge in unregulated hunters
484 D.S. Hammond

and wildlife collectors, illicit smugglers, scape-scale conservation gaps based on this
socially transmitted diseases (e.g. Palmer et synthesis. The chapter also provides a brief
al., 2002), pollution (see Chapter 8), overview of approaches assessed and
cultural transformation and social upheaval employed in the quest for sustainable forest
to regions previously insulated from the management, including reduced impact
widespread effects of frontier natural logging techniques, extractive reserves and
resource use by the constraints on river concession and community-based manage-
navigability. ment. It explores some of the key challenges
in achieving adequate conservation and
habitat protection across the region and
Isolation those facing efforts to sustainably utilize
various forest resources towards further
From many perspectives, the Precambrian economic development. The chapter con-
geology of the Guiana Shield has been iso- cludes with a perspective on forest manage-
lating. Whether viewed as isolation from ment that highlights briefly how lack of
the sediments of the Andes, from the fertil- individual opportunity and a mismatch of
ity of the varzea, through the geo- objectives with biophysical processes can
logical control on river navigability or from limit success in achieving sustainable man-
the sociocultural complexity that agement.
appears to have and continues to typify
the shield perimeter, the Guiana Shield
interior appears as an isolating environ- Conservation Patterns and Approaches
ment. It is precisely this long-standing iso-
lation that characterizes the relatively Areas designated for forest protection
higher risk attached to uncontrolled and
poorly managed forestland use in the The Guiana Shield contains a world-class
Guiana Shield compared to other regions system of protected areas that has grown
arguably more strongly shaped by prehis- from one of the first areas officially desig-
toric human occupation and catastrophic nated for protection in South America
events. (Kaieteur National Park, Guyana in 1929)
through to the largest protected area ever
established in the neotropics in 2002
Modern forest cover and deforestation rates (Tumucumaque National Park, Brazil)
(Table 9.2). The region boasts three major
Perhaps as a consequence of these factors, World Heritage Sites, including the Central
even today deforestation rates and forest Suriname Reserve (designated 2000), the
area coverages remain some of the lowest Central Amazon Conservation Complex1 in
and highest in the world, forming a signifi- Brazil (2003) and the oldest, Canaima
cant and growing part of the standing trop- National Park in eastern Venezuela (1994).
ical forest area in South America and the Three coastal areas have also been desig-
world (Table 9.1). As deforestation trends nated as Wetlands of International
continue elsewhere, the Guiana Shield will Importance (Ramsar), two at Basse-Mana
increasingly represent a greater share of the (established 1993) and Les Marais de Kaw
remaining closed forest cover, unless land- (1993) along the French Guiana coast and
use practices and patterns change dramati- another covering the tropical mangrove sys-
cally. tem at Coppenamemonding in Suriname
The purpose of this chapter is to review since 1985.
and synthesize available information on the The rate of growth in area allocated to
conservation and management of the shield protection in the Guiana Shield has
region. It characterizes the history, effort matched or exceeded the accrual rate glob-
and modern approaches to habitat protec- ally (Fig. 9.1). The establishment of new
tion in the region and offers a view on land- protected areas around the world has
Table 9.1. General status of tropical forest cover in the Guiana Shield in relation to continental and global standing area. Source: Land and total forest area –
FAO (2000); land area in Guiana Shield – OCEI/MARNR (2000), IBGE (2000), FAO (2000); tropical forest area in Guiana Shield – FAO (2000), except Venezuela
based on Huber (1995) and Brazil based on IBGE municipio estimates.

Land Area (km2) Intact Forest Area (km2) Deforestation (%)


Total in GS Total Tropical Tropical in GS 1990–2000

French Guiana 88,150 88,150 79,260 29,260 79,260 –


Suriname 156,000 156,000 141,130 141,130 141,130 –

Forest Conservation and Management


Guyana 214,980 214,980 168,790 163,726 163,726 2.8
Venezuela 882,060 453,950 495,060 346,542 287,630 4.2
Brazil 8,456,510 1,204,279 5,439,050 4,133,678 945,573 4.1
Colombia 1,038,710 170,500 490,601 412,105 170,500 3.7
Guiana Shield 10,836,410 2,287,859 6,813,891 1,342,246
South America 17,547,510 8,856,180 5,490,832 26.9%
World 148,000,000 38,694,550 10,060,583 13.2%

485
486 D.S. Hammond

Table 9.2. The 35 largest protected areas legislatively established in the five countries of the Guiana
Shield. Only publicly owned areas restricted to low-impact use are included. IUCN category: Ia –
Strict Nature Reserve managed for science, Ib – Wilderness Area managed mainly for wilderness pro-
tection, II – National Park managed for ecosystem protection and recreation, III - Natural Monument
managed for conservation of specific natural features, IV – Habitat/Species Management Area man-
aged for conservation through management intervention.

Rank Protected area Designation Country IUCN Year Est Area (km2)

1 Montanhas de Tumucumaque National Park Brazil II 2002 38,670


2 Formaciones de Tepuyes Natural Monument Venezuela III 1990 34,200
3 Parima-Tapirapecó National Park Venezuela II 1991 33,100
4 Canaima National Park Venezuela II 1962 30,000
5 Pico da Neblina National Park Brazil II 1979 22,000
6 Jau National Park Brazil II 1980 22,720
7 Central Suriname Reserve Nature Reserve Suriname II 1998 16,000
8 Serrania La Neblina National Park Venezuela II 1978 13,600
9 Serrania de Chiribiquete National Park Colombia II 1989 12,800
10 Puianawai Nature Reserve Colombia III 1989 10,920
11 Nukak Nature Reserve Colombia III 1989 8,550
12 Cavo Orange National Park Brazil II 1980 6,190
13 Delta de Orinoco National Park Venezuela II 1991 5,698
14 Uatuma Biological Reserve Brazil Ia 1990 5,600
15 Viruá National Park Brazil II 1998 2,270
16 Rio Treombetas Biological Reserve Brazil Ia 1979 3,850
17 Lago Piratuba Biological Reserve Brazil Ib 1980 3,570
18 Anavilhanas Ecological Station Brazil Ia 1981 3,500
19 Jau Sarisariñama National Park Venezuela II 1978 3,300
20 Yapacana National Park Venezuela II 1978 3,200
21 Niquia Ecological Station Brazil Ia 1985 2,866
22 Jari Ecological Station Brazil Ia 1982 2,271
23 Serra da Mocidade National Park Brazil II 1998 3,610
24 Duida-Marahuaca National Park Venezuela II 1978 2,100
25 Iwokrama Wilderness Preserve Guyana Ib 1997 1,800
26 Monte Roraima National Park Brazil II 1989 1,160
27 Maracá Ecological Station Brazil Ia 1981 1,013
28 Sipaliwini Nature Reserve Suriname IV 1972 1,000
29 Nouragues Nature Reserve French Guiana Ia 1995 1,000
30 Marais de Kaw-Roura Nature Reserve French Guiana IV 1998 947
31 Caracarai Ecological Station Brazil Ia 1982 806
32 La Trinité Nature Reserve French Guiana Ia 1996 760
33 Maracá-Jipioca Ecological Station Brazil Ia 1981 720
34 Kaieteur (extended) National Park Guyana II 1929 (98) 630
35 Forêt de Saül Prevectorially Decreed French Guiana IV 1995 600
Biotope

tapered off since 1994, after experiencing a erably by country, but several distinct pat-
22-year span of consistent annual increase terns are discernible. First, the rates of
in area since 1972. The establishment of countries with only a part of their national
Canaima National Park in 1962 represented territory in the region, viz. Colombia,
one of the earliest efforts to protect large Venezuela and Brazil, is considerably
areas of tropical forest in the world and the higher than those with their entire (depart-
most significant contiguous area designated mental) area in the region, i.e. French
in the Guiana Shield until the 2002 estab- Guiana, Suriname and Guyana (Fig. 9.2).
lishment of the Montanhas de The difference is logical, given the need for
Tumucumaque (Table 9.2). The annualized the smaller, shield-bound countries to con-
rate of protection since national system sider all of their potential land-use options,
establishment (in the region) varies consid- while their larger neighbours have other
Forest Conservation and Management 487

territory to consider in meeting their since the early 1990s (Ramdass and Haniff,
national economic goals. Nonetheless, the 1990; Agriconsulting, 1993). Hopefully this
rate of protected area accrual in Guyana will alter the country’s current standing as
remains well below that achieved in the smallest contributor to the regional net-
Suriname or French Guiana since establish- work of protected areas.
ment of the inaugural protected area in each By 2003, this network amounted to
country. Guyana’s rate of accrual is clearly 1.6% of the global area protected, based on
affected by the very early establishment of IUCN-WCMC calculations of 18.763 million
Kaieteur National Park, but even if adjusted km2 of area officially designated worldwide
to the year of independence in both Guyana as protected. This equates to 4.6% of IUCN
(1966) and Suriname (1975), Guyana’s rate Category I–III areas registered globally by
remains much lower than the 400 km2 per 2003 (Chape et al., 2003). Nearly 13% of
annum that would place it in line with Guiana Shield forests are estimated to be
other countries in the region. The develop- under formal protection through six
ment of a national protected area system, national systems, although the proportion
after having legislated for the 1600 km2 of each country’s territory under protection
Iwokrama Forest Wilderness Preserve in in the region varies from just over 1% in
1997, was imminent by 2003 and has been Guyana to nearly 28% of Venezuelan
forming around a series of designated sites Guayana (Table 9.3). On average, there are

Global
% total allocated area

Guiana
Shield

Fig. 9.1. Twentieth-century growth trends in habitat protection worldwide and in the Guiana Shield
depicted as percentage of all officially recognized protected area in 2002. Sources: global (http://sea.unep-
wcmc.org/wdbpa, but see Chape et al., 2003), Guiana Shield (http://sea.unep-wcmc.org/wdbpa cross-ref-
erenced with size and establishment data from national park and protected area agencies and
administrators – INPARQUES, IBAMA, EPA-Guyana, INDERENA). Figures do not include areas proposed
but without legislative mandate, designated Biosphere Reserves (except where these include national sys-
tem units), or Amerindian reserves, resguardos, titled or ancestral land areas (see Chapter 8).
488 D.S. Hammond

Annualized rate of protection (km2/ yr ) 10000


Brazil
Venezuela
Colombia

1000
Suriname

French
100 Guiana

Guyana

10
10,000 100,000 1,000,000 10,000,000
2
Land area in Guiana Shield (km )

Fig. 9.2. Relationship between rate of area protection (log10) and total area (log10) within Guiana Shield
for each of the six countries forming the region since inception of their respective protected area systems.

nearly 7.5 ha of forest under protection for concessions (e.g. biological reserves in tim-
every person living in the shield region, ber concessions, Guyana). Of course, sim-
based on 2000 population estimates (Table ple measures of proportional forest area
9.3). receiving legal protection do not adequately
The nearly 290,000 km2 (almost the address whether the spatial distribution of
size of Guyana and French Guiana com- protected area: (i) captures the main fea-
bined) of protected area already established tures of the forest landscape; (ii) is likely to
across the region covers an estimated retain long-term conservation value; or (iii)
12.3% of tropical humid forest area remain- is functionally performing according to the
ing globally (Chape et al., 2003). This is a conservation objectives established through
substantial figure and likely to rise when legislative mandate. It does, however, iden-
processes in Guyana and French Guiana tify a continuing commitment to forest pro-
lead to the formal expansion of their tection in the region that exceeds that
national (departmental) systems of protec- typically encountered in most other regions
tion that currently lag other regional com- of the world.
mitments (Table 9.3). At the same time
tropical wet forest area elsewhere continues
to decline at a faster pace, pushing upwards Landscape conservation assessments
even further the fractional contribution of
Guiana Shield areas to global protection of Several large global conservation and envir-
this general forest biome. onmental organizations have increasingly
The vast area under protection within taken on the self-designated task of assess-
the shield is distributed over 51 distinct ing and classifying the terrestrial and
units exceeding 10 km2 (35 largest in Table marine regions of the planet according to
9.1) (Fig. 9.3). Numerous units with an area their conservation value, immediacy of the
less than 10 km2 also contribute to forest threats confronting their persistence and
protection and research under systems of priorities for investment in their protection.
ecological reserves established directly by The Guiana Shield has formed part of the
national forestry services or as part of area assessed through these approaches.
required management planning in timber The results and relevance of these global
Table 9.3. General status of habitat protection in the Guiana Shield.

Venezuela Colombia Guyana Brazil Suriname French Guiana Guiana Shield

Total area in GS (km2) 453,950 170,500 214,980 1,204,279 156,000 88,150 2,287,859

Forest Conservation and Management


Population (yr 2000) 1,544,915 53,650 697,286 1,041,595 431,303 157,213 3,925,962
Total PA (km2) 125,243 32,270 2594 120,816 16,359 4970 302,252
% of GS area in PAs 27.6 18.9 1.2 10.0 10.5 5.6 13.2
Year initial PA established 1962 1989 1929 1979 1961 1989
Protected area per person (ha) 8.1 60.1 0.4 10.4 3.8 3.2 7.4

489
490 D.S. Hammond

noco
O ri

Ne
gr o

n
az o
Am

13

noco
Ori

19
2 4 26 34
25 32 30
20 7 12
27 29
24
11 10 3 33
31 2 1
21 15
8 15
9 23
5 Equator

Ne
16
gr o 22
14
6 18 az
on
Am

Fig. 9.3. Spatial distribution of official protected area (solid black) across the Guiana Shield and wider
neotropics (inset). Numbered units are consistent with rank in Table 9.2. Designated Biosphere Reserves
extending beyond boundaries of protected areas (right diagonal hatching) and proposed conservation units
in Guyana (left diagonal hatching) are also presented. Forest cover is solid grey. Spatial data sources:
Brazil: IBAMA; Venezuela: INPARQUES and WCPA/WCMC; Guyana: Iwokrama, CI-Guyana, EPA;
Suriname: CELOS; French Guiana and Colombia: WCPA/WCMC.

initiatives to shield features are summa- to the shield area (as defined in this vol-
rized here. ume) (Fig. 9.4). The four largest ecoregions
effectively dissect the shield area into quad-
Ecoregions rants that embrace an archipelago formed
from the three smaller, edaphic-based
Olson et al. (2001) fractionated the plane- units. Other smaller units largely reflect
tary terrestrial surface area into 867 eco- variation in coastal vegetation between the
regions based on a combination of Orinoco delta and Amazon mouth and
landforms, vegetation types and variation along the upper Orinoco, Solimões and Rio
in ecological processes.2 Negro.
As part of this global classification sys- While savanna, campinarana and tepui
tem, the Guiana Shield is dissected into 14 formations define distinct, often abrupt,
ecoregions, the seven largest being endemic edaphic transitions, the four larger eco-
Forest Conservation and Management 491

regions cover more territory and embrace region into at least three distinct ecoregions
considerably greater spatial variation in for- (dashed lines in GMF ecoregion of Fig. 9.4).
est composition. Forest-type classifications The Guayana Highland ecoregional area is
in Guyana and the Guianas comprehen- consistent with geological transitions asso-
sively analysed by ter Steege (1998) and ter ciated with distributions of Roraima sedi-
Steege and Zondervan (2000) more coher- mentaries, Uatuma volcanics and Parguaza
ently delimit compositional transitions that granites that dominate Venezuelan
are not consistent with margins of the Guayana, although east to west precipita-
Guiana Moist Forest ecoregion. This dis- tion gradients associated with the Savanna
crepancy can for the most part be attributed Trough are not fully delimited. The change
to geology. Transitions between areas of from this unit to the adjacent Negro-Branco
exposed crystalline basement dominated by Moist Forest unit overlying the Casiquiare
TATE granitoids, greenstone belts and Rift reflects an important landscape forest
Phanerozoic sands (of the Berbice transition. Perhaps more importantly, the
Formation) (see ‘Greenstone belts’, Chapter geology also spatially defines the largest
2) provide a spatial delimiter of forest type and growing threat to forest integrity and
associations that dissect Olson et al.’s function: unregulated mining.

Fig. 9.4. Seven major ecoregions unique to the Guiana Shield (after Olson et al., 2001). GMF: Guianan
moist forests; U-T MF: Uatuma-Trombetas moist forests; GS: Guyanan savannas; GH: Guayanan highlands
moist forest; T: Tepuis; RNC: Rio Negro campinarana; J-S-N MF: Japurá-Solimões-Negro moist forest.
Ecoregion spatial coverage: WWF-USA. Thick line demarcates southern boundary of Berbice Formation, a
unique ecoregion candidate currently subsumed within the GMF.
492 D.S. Hammond

Forest frontiers water quality and sedimentation and is


widely implemented as part of sound
A four-point classification of frontier forest forestry practices globally (Dykstra and
threat status by Bryant et al. (1997) identi- Heinrich, 1996). Yet, placer gold deposits
fies forests under medium or high threat of are located in these very same creeks and
degradation along most of the perimeter rivers making them subject to massive sedi-
area of the Guiana Shield (Fig. 9.5A). Three ment influxes as a consequence of current
core areas of low threat were assigned to the extraction techniques (see ‘Commercial
main upland ‘islands’ forming the Guiana mining’, Chapter 8), considerably reducing
Shield, i.e. the Tumucumaque Uplands, the functional value of buffer zoning.
Guayana Highlands and Chiribiquete
Plateau (and eastern Colombian lowlands)
Hotspots
(see ‘Shield macro-features’, Chapter 2).
The distribution of threat magnitude Areas of high conservation value that are
assigned by Bryant et al. appears largely disproportionately threatened with loss or
linked to areas allocated to selective logging degradation have been variously classified
(Fig. 8.17) and, to a lesser degree, agricul- as biodiversity hotspots (Davis et al., 1997;
ture. Low threat status is widely assigned to Myers et al., 2000). None of these is located
major mining regions containing some of in the Guiana Shield region. Areas that
the largest greenstone belts in the world, have extensive forest cover and face few
including areas in Suriname, French imminent threats are collectively included
Guiana and Amapá (Fig. 9.5B). Mining as a Wilderness Area. Virtually the entire
activity throughout the valleys and along Amazon Downwarp, Sub-Andean Foredeep
the periphery of the Roraima sedimentaries and Guiana Shield comprise the largest
in Venezuelan Guayana also suggests that tropical Wilderness Area identified.
assignment of low risk status is not appro- Hotspots are delimited by the esti-
priately weighting the history and future of mated fraction of ‘original’ vegetation cover
unregulated mining activity. Both geologi- lost combined with the estimated fraction
cal groups are strongly associated with gold of endemics composing this original cover.
and diamond-bearing substrates that have While relatively little forest cover has been
attracted vast numbers of small-scale min- lost across the Guiana Shield, several
ers seeking subsistence incomes (see important features suggest that an assess-
‘Commercial mining’, Chapter 8). ment of this kind based on recent popula-
The pressures to keep valuable green- tion growth and deforestation trends alone
stone formations open to mining are high may not sufficiently embrace future vulner-
and areas that have already been subjected abilities. The low-energy attributes that
to mining are not normally considered for characterize modern forest processes in the
conservation. It is not surprising, therefore, Guiana Shield (see Chapters 2, 3 and 7) sug-
that countries with large greenstone belts gest that this region’s susceptibility to cata-
and a large number of registered mines strophic change is disproportionately
adjoining these areas have relatively little higher than other regions of the neotropics.
of this area allocated for habitat protection The large number of endemic plants in rel-
(Fig. 9.6) and widely overlap with areas atively high local abundances but restricted
allocated for timber production (Fig. 9.7), geographic ranges within the shield region
although these two resource-use practices (ter Steege, 2000) also argues for a rethink
and the regulations designed to moderate on how methods for assessing conservation
their impacts are often highly incompatible. value cope with geomorphography and its
For example, the Guyana Forestry influence on the way climate, substrate,
Commission’s Code-of-Practice establishes phylogeny and people combine to influence
a minimum creek buffer zone applicable in forests across these regions. Across the
all commercial forestry operations. This neotropics, geomorphographic control on
sensibly mitigates the impact of logging on forest processes renders the Guiana Shield
Forest Conservation and Management 493

zon

oco
Medium-high
Orin

Low
Low

Low Low

Ne
g ro

on
az
Am

Unassessed Medium-high

Fig. 9.5. Two-point spatial threat assessment of regions within the Guiana Shield (after Bryant et al., 1997).
Threat assessment spatial coverage: WRI.
494 D.S. Hammond

Fig. 9.6. Greenstone belts, gold mines and protected areas. Percentage of land area covered by green-
stones, the percentage of registered gold mines located within 10 km of the nearest greenstone formation
and percentage of these greenstones located within existing protected areas (IUCN I–IV categories) (see
Fig. 9.2). Greenstone distribution based on Gibbs and Barron (1993). Mine locations based on USGS and
CVGTM (1993).

particularly vulnerable to widespread provenance. Rainfall and temperature


change at comparably modest levels of variation, in relation to topography and
human intervention, particularly when this geographic location, exert a larger control
is poorly managed. over soil–vegetation relationships (see
‘Soils and soil fertility’ and ‘Climate and
weather sections’, Chapter 2), as may his-
Habitat and forest types as conservation torical patterns of human influence (see
units Chapter 8).
Dissecting the Guiana Shield by the
Recognizing aquatic habitat groups based spatial distribution of these aquatic and ter-
on salinity and dissolved organic carbon restrial habitats arguably defines an opti-
content helps to separate marine and estu- mum scale for discriminating distributions
arine habitats along the Atlantic coast from of all but the largest (e.g. jaguar, puma,
freshwater systems characterized by widely tapir) or most mobile (e.g. migrants) of
varying DOC, TSS and TZ+ (Furch, 1984) species (see Chapter 4). There are several
(see ‘Hydrology’, Chapter 2). Discri- reasons for this optimum.
mination of terrestrial habitats according to First, geomorphographic control delim-
vegetation type strongly shadows variation its, at the widest scale, the range of possible
in edaphic attributes (Davis and Richards, edaphic and aquatic attributes. For exam-
1933; Fanshawe, 1952; Richards, 1952; ple, the absence of significant sources of
Schulz, 1960; Heyligers, 1963; Cooper, mineral calcium in the shield, and thus cer-
1979; Lescure and Boulet, 1985; ter Steege tain edaphic and aquatic conditions,
et al., 1993; Duivenvoorden and Lips, 1995; reflects both the geographic position (rain-
Coomes and Grubb, 1996; ter Steege, 2000; fall and temperature patterns) and weather-
ter Steege and Hammond, 2001), par- ing age of the region.
ticularly in relation to surface and soil Secondly, the underlying Precambrian
moisture status, but also parent material geomorphology exerts a pronounced influ-
Forest Conservation and Management 495

ence over drainage dynamics and the extent late Cretaceous and prior to the uplift of
and magnitude of hydrological disturbance both the Andean mountains and the
over the lowland edaphic environment. Panamanian land bridge, terrestrial and
The low-energy system characterizing the freshwater life were already evolving on a
surface drainages of the shield region pro- shield landscape that was dominated by
mote autochthonous over allochthonous slow gradation and weak diastrophism. In
pathways to soil development. Thus, long- comparison, the western South American
standing differences in parent material and environment was believed to be largely epi-
climate resonate more significantly through continental (shallow marine) (see Chapters
the soil development process. In regions 2 and 7). This would suggest a much longer
where widespread import of externally legacy of continuous in situ terrestrial and
sourced materials is delivered through rela- freshwater evolution across the major
tively high-energy fluvial systems, more fre- shields relative to other regions of the
quent soil turnover would reduce edaphic neotropics. While proxy measures suggest
patchiness and increase variation at much climate fluctuated significantly over at least
smaller mixing scales. the last 80 million years, no evidence cur-
Thirdly, habitats in the Guiana Shield rently points to a wholesale extinction
represent one of the oldest existing terres- across the Guiana Shield. This cannot be
trial regions in the neotropics. During the discounted, but neither can it be conclu-

Fig. 9.7. Zones of overlapping mining (empty circles) and timber production (black) across the north
Guiana Shield. See Figs 8.8 and 8.17 for data sources.
496 D.S. Hammond

sively supported from palaeontological evi- region are typified by a very high abun-
dence derived from other parts of the dance to range ratio (e.g. Alexa spp.,
neotropics and then scaled up across very Dicymbe spp., Eperua spp., Chlorocardium
large regions that have contributed few, if rodiei, Dicorynia guianensis, Catostemma
any, fossilized remains. spp., Micrandra spp., Mora gonggrijpii).
While some lineages undoubtedly dis- Common species in Central America and
appeared from the region during western Amazon appear to have much
unfavourable shifts in climate, it is difficult larger geographic ranges with much lower
to counter a much earlier start to terrestrial relative abundances (e.g. Iriartea deltoidea,
evolution of life in the shield. Given the rel- Poulsenia armata, Pseudomeldia spp.).
ative stability of the region’s geographic Uniqueness, an important component of
position over most of the Cenozoic, it conservation value, thus equates differently
would have supported the accumulation in the Guiana Shield than it does in other
of biotic attributes that bring a compara- neotropical regions affected by different
tive advantage in low-energy environments. geomorphographic controls. These regions
The primacy of meso-scale edaphic spe- are also likely to express components of
cialization in defining opportunity in conservation value differently.
this competitive environment trans- In the Guiana Shield, delineation
lates into a large number of modern of conservation units based on forest
forest types dominated by one or few type characterization creates greater oppor-
endemic species that are hyper-abundant tunities for successfully meeting the chal-
within their highly restricted geographic lenges that confront representative
ranges and express attributes that place a conservation. Unique habitats or forest
premium on high per capita survival, not types across the northern shield region
dispersal. work well as working units because these
At larger classificatory scales, much of are strongly linked to transitions between
this important habitat variation is lost. relatively large-seeded endemics with rap-
Clear distinctions identified by foresters, idly diminishing representation at larger
ecologists, botanists and natural historians scales. Refined soil-type classification
over the last century or more are merged cross-referenced with geographic ranging of
into single associations. The drawback at endemics, in this instance, would yield an
these large scales is that the aggregation of optimum conservation topology if the
wide-ranging patchiness simply reinforces objective is representative forest system
notions of tropical environmental unifor- protection.
mity. Ironically, viewing systems at this
large scale supports the primacy of smaller-
scale processes, that strongly influence Management Approaches and their
alpha-diversity levels (e.g. Hubbell et al., Applicability
1999), as the nexus of conservation deci-
sion-making. The important role of the Mainly protection
physical environment in structuring varia-
tion at meso-scales is removed. Meso-scale, As a management approach, protected areas
habitat or beta-diversity, is lost in the analy- (after IUCN categories I–III) aim to carry out
sis. a series of functionally important roles, at
Across much of the Guiana Shield least in theory. According to the IUCN, pro-
where alpha diversity is relatively low (ter tected forest areas should: (i) maintain for-
Steege et al., 2000), and large populations est ecological and genetic processes; (ii)
of narrow-ranging endemics are common minimize artificial disturbances; (iii) pro-
(Fanshawe, 1952; Richards, 1996), the col- vide opportunities for low-impact research,
lapse of habitat variation into larger units education and recreation; and (iv) protect
fails to emphasize the value of this uncom- outstanding natural features and scenic
mon feature. Endemic species in the shield areas of national or international signifi-
Forest Conservation and Management 497

cance. In practice, several of these broad- 1. Alter the standing forest stock of life-his-
based, generic objectives can prove difficult tory attributes (Hammond and Brown,
to achieve without substantial tailoring at 1995; ter Steege and Hammond, 2001);
the conservation unit level for a number of 2. Affect how these bring comparative
reasons. advantage to different plant and animal
taxa under prevailing conditions;
Ecological and genetic processes 3. Shape the relative influence of determin-
istic (e.g. historical human forest use) and
One difficulty is defining measurable indi- stochastic (e.g. density-dependent mortal-
cators of forest ecological and genetic ity) processes on forest change; and
health that are appropriate to the system 4. Regulate how fast the balance of standing
being protected. For example, protection of taxa and biomass is changing as a conse-
forests in the western United States over the quence.
last half century has implicitly involved
thwarting significant disturbance from fire, ‘Artificial’ disturbances
largely as a consequence of the catastrophic
forest and human loss caused by the Great Consequently, separating disturbances that
Fires of 1910 (Pyne, 2001). Massive, stand- form part of the ‘natural’ ecological process
replacement fires at the turn of the 21st cen- from those that are ‘artificial’ can also prove
tury have now prompted a re-think about difficult, if forests being protected prove to
how best to define ‘natural’ processes and have been subject to widespread, long-last-
how best to ‘protect’ these. Fires are seen ing and/or intensive prehistoric use. Many
now as an inimical part of most forest eco- small protected areas in Central America
logical processes and subduing their influ- previously believed to have been ‘pristine’
ence over decadal scales has created have subsequently proven affected in part
conditions literally fuelling catastrophic by pre-Columbian agricultural use (e.g.
stand replacement events at landscape Barro Colorado Island, Panama: Piperno,
scales. 1990; La Selva, Costa Rica: Kennedy and
Across the neotropics, regional forest Horn, 1997) as have many protected forest
systems are also subject to different large- areas in the Guiana Shield (e.g. Kanuku
scale influences on local ecological Mountains, Guyana (proposed): Evans and
processes. While these invariably overlap, Meggers, 1960; e.g. Iwokrama, Guyana:
the Guiana Shield area is clearly exposed to Williams, 1994; Nouragues, French Guiana:
far fewer of these (Fig. 7.2; see Chapters 2 Ledru et al., 1997). Again, the problem
and 7). Charcoal and historical evidence arises in discriminating the relative effects
point to fire as a widespread natural agent of these chronological series of events from
of catastrophic change across the shield other stochastic processes at work. If the
region, other parts of the eastern Amazon accumulated archaeological evidence
and, possibly, parts of Panama and Central reflects the true distribution of prehistoric
America. In contrast, the western Amazon human impacts, then chronologies of
has yet to be characterized as a fire-affected human impact across the interior of the
system. Rather, it appears more strongly Guiana Shield should be compressed rela-
controlled by high-energy hydrological dis- tive to Central America, the Amazon
turbances (Räsänen et al., 1987) (see Downwarp and Sub-Andean Foredeep
Chapters 2 and 7). Equally, the relative regions.
impact of pre-Columbian indigenous soci-
eties appears to vary spatially (see Chapter Research, education and recreation
8). The relative impacts of these different
trajectories on smaller-scale ecological Unlike the more diffuse realities of environ-
processes are not readily thrown under the mental protection, these real-time activities
same definition or monitored using the are more amenable to practical manage-
same criteria because they may: ment. The main difficulty in generalizing
498 D.S. Hammond

arises with the thresholding definition of communities to the same or greater degree
low-impact. Take the example of fire. than has been done at international levels.
Across many protected areas in the Significant areas across the shield
neotropics the creation of managed fire region have been established for habitat
treatments as part of research would not be protection and this coverage already rivals
considered among the low-impact, allow- all other tropical forest regions for the title
able uses. Yet, ecotonal forests in some pro- of ‘best protected’. Only Guyana (and to a
tected areas along the perimeter of tropical lesser extent French Guiana) has yet to legi-
savannas are frequently affected by fire slate for areas that have been under consid-
(Thompson et al., 1992). If the objective of eration since the 1990s, but this is likely to
restricting use is to minimize human influ- change. The region boasts nearly 30% of
ence over ‘natural’ system processes, then remaining closed tropical forest cover in
the allowable range of activities associated South America and 13% of that estimated
with these general forms of low-impact use to occur globally (Tables 9.1 and 9.2).
are again subject to local considerations of Approximately 20% of lowland tropical
suitability. In the Guiana Shield many loca- forest in the region is protected and this is
tions would require significantly lower likely to rise to 25% by 2010.
impact allowances than parallel communi- With one-quarter of the area protected
ties within ramping system regions. and all countries in the region already or
expected to have more than 10% of their
land area committed to strict habitat protec-
Protected area management in the Guiana tion, tailoring objectives and identifying
Shield practices that will lead to relevant conser-
vation in these areas would appear crucial.
Strategies and methods for prioritizing and Moreover, articulating these through well-
selecting protected areas across the Guiana considered management plans and coordi-
Shield have dominated processes associ- nated actions has remained noticeably
ated with protected area management absent from many designated areas (e.g.
(Hoosein, 1996; Rodríquez and Rojas- Huber, 1995, 2001). In part this is due to a
Suárez, 1996; Huber, 1997; Stattersfield et shortfall in resources needed to effectively
al., 1998; ter Steege, 1998; Funk et al., 1999; manage for conservation in a region being
ter Steege, 2000; Mittermeier et al., 2001; increasingly used for its mineral, timber
Funk and Richardson, 2002). Exploring and wildlife resources and weighed down
appropriate objectives and methods of man- by excess external debt burdens, poor
aging established areas has received less access to education and health care, bouts
region-wide attention, although this is of social unrest, high rates of emigration
arguably the major constraint facing func- and mounting costs of infrastructural main-
tional conservation in most areas. Bruner et tenance (see Chapter 8).
al. (2001) concluded that the number of
park guards was the best determinant of Mainly timber
protection effectiveness in tropical parks
based on questionnaire surveys of park Management for timber production has
managers and park staff (56%), NGOs and been a long-standing focus across many
researchers (30%) and protected area agen- regions of the Guiana Shield since the early
cies (14%).3 Defending protected areas 1900s. Regional notions of sustainability,
through enforcement and patrolling invari- however, have broadened beyond mainte-
ably plays an important role in deterring nance of harvestable volume to include
illegal activity. But taken as a top priority, many non-commodity benefits (e.g. formu-
enforcement and patrolling is a blunt lated in the Tarapoto Agreement).
instrument that is symptomatic of failure, Nonetheless, timber remains the largest for-
rather than success, in effectively demon- est sector contribution to national GDP,
strating protected area benefits to local apart from mining, and efforts to improve
Forest Conservation and Management 499

timber production and its management addressing these shortfalls and refining and
have explored several avenues (e.g. King, redirecting pre-site management effort
1963). These have focused principally on (McNabb and Wadouski, 1999), but nutri-
plantation and natural forest management ent emigration through biomass removal,
approaches to timber production and the particularly scarce calcium, continues to
potential for supplying material to paper proceed at very high rates in stands of
pulp, (peeled and sliced) veneer, plywood, Eucalyptus urograndis (Spangenberg et al.,
sawn split and hewn wood markets (e.g. 1996).
Vink, 1970; Welch, 1975a). By the close of the 20th century, very
few plantations were operating within the
Plantation forestry Guiana Shield region. Although Brazil pro-
duced between 32 and 46 million m3 of
The prospect of meeting domestic demand plantation wood fibre alone, a mere 4–5%
and supplying export markets with timber of this originated in the Guiana Shield and
grown in a plantation setting has largely virtually all of this was associated with the
proven unworkable over most parts of the Jari pulpwood operation (IBGE, 2002). As of
Guiana Shield where trials have been con- 2000, no other commercial-scale planta-
ducted (Vink, 1970; de Graaf, 1986). A wide tions are producing significant quantities in
range of native commercial timber species the region.
have been examined, including Simaruba, Plantations hold tremendous prospect
Carapa, Cedrela, Peltogyne, Dipteryx, in providing a reliable supply of light hard-
Centrolobium, Caryocar, Anacardium, wood and wood fibre to supply local down-
Virola, and Dicorynia with little success stream industries and market demand (F.
(British Guiana Forest Department, Wadsworth, personal communication).
1940–1965; Welch, 1975a; de Graaf, 1986). Little success has been met across the
The potential of faster-growing exotics such Guiana Shield in establishing commercially
as Pinus caribaea, Eucalyptus spp. and viable plantations of both natural and
Gmelina arborea, as well as valuable hard- exotic species. The principal difficulty aris-
woods, such as Swietenia, Khaya and ing is the cost-effectiveness of producing
Tectona was also explored with mixed suc- wood through this approach. Relatively
cess due to the high costs of pest and weed slow growth increments combined with
management combined with disappointing elevated pest and weed management costs
growth increments (Vink, 1970). For exam- and stagnant prices for low-end tropical
ple, one trial of P. caribaea grown on ferra- timber and fibre work to reduce or elimi-
solic soils in Guyana achieved an annual nate profit margins. Improving growth,
average increment of 1.8 m3/ha over a 20- cost-effectiveness of management or timber
year period, a figure only marginally prices would potentially make plantation
improving on estimated volume accrual in forestry a more attractive investment. In
many natural forests of the region. Perhaps many parts of the Guiana Shield where
the most outstanding example of plantation forests have been degraded through high-
failure is the case of the pulpwood opera- frequency selective logging, plantations
tion established along the lower Jari River under these circumstances might offer a
in south Amapá/east Pará states (Rollet, means of improving landscape fertility
1980; Palmer, 1986). Monocultured (Lugo, 1995), increasing wood supply and
Gmelina and Pinus were battered by fungal offering greater flexibility in allocating nat-
and insect damage and rapid declines in ural forest areas to lower-impact direct and
productivity, principally through disregard indirect uses. It is unlikely, however, that
for the role of internal nutrient cycling the accelerated growth required from plan-
(Russell, 1987) and density-dependent tation-grown timber would foster the same
attack in maintaining productivity. Over strength and durability properties (see
the last 20 years, however, some productiv- ‘Wood density’, Chapter 7) that have typi-
ity gains have been made, primarily by fied the main line of exported timber prod-
500 D.S. Hammond

ucts from the Guiana Shield for centuries planning of felling and extraction (e.g.
(e.g. Mackay, 1926). Hammond et al., 2000). As a consequence,
operational efficiency is improved and
Natural forest management wood waste reduced in comparison with
unplanned approaches (Boltz et al., 2003).
Managing standing mixed forests of the Yet, many operators have been slow to
interior for long-term production of these implement these practices, in part because
heavy hardwood products holds consider- the management of more complex
ably greater promise if natural regeneration approaches to harvesting bear with them
is employed as the main avenue to restock- additional costs beyond forest operational
ing. Systematic selective logging has been considerations (Hammond et al., 2000). The
the main route to timber production in the financial benefits of employing RIL tech-
Guianas since the early 20th century niques are often slim, neutral or slightly
(McTurk, 1882; Hohenkerk, 1922), although negative (Barreto et al., 1988; Winkler,
pro-active effort to manage forests for tim- 1997; van der Hout, 1999, 2000; Armstrong
ber was not taken up until the 1950s. and Inglis, 2000), suggesting even short-
A number of silvicultural approaches term lapses in management rigour could
have been explored in an attempt to direct create further financial losses. These ‘other’
natural regeneration towards greater tim- costs become even more significant when it
ber-tree stocking. Among the most widely is apparent that they may not lead to any
known are thinning (e.g. through poisoning immediate financial benefits from a market
and girdling) (King, 1965; Jonkers and that is slow to pay higher tropical timber
Schmidt, 1984; de Graaf et al., 1999), prices (Barbier et al., 1994), despite efforts
enrichment planting, climber and liana to structure incentives for better forest
elimination (e.g. Putz, 1991), seed tree stewardship through certification
retention (e.g. Plumptre, 1995) and stand approaches. Operators are largely unwilling
damage control (e.g. van der Hout, 2000). to bear greater up-front costs and greater
Combined, these techniques form a process financial risk in the face of longer-term
that aims to modify over time the existing political instability and uncertain land
mix of species and size classes in a way that tenure (Pearce et al., 2003). In effect, grow-
improves timber value and manageability ing and harvesting trees through natural
of the stand, often referred to as ‘domestica- forest management approaches in the trop-
tion’ of natural forest (Dickinson et al., ics requires adoption of an investment hori-
1996; de Graaf, 2000). Employed à la carte, zon that extends well beyond the limits of
they can act as effective tools in ensuring normal risk tolerance. Across the Guiana
that forest functioning remains largely Shield, this horizon may extend beyond the
unimpeded, although (at least transitory) lifetime of the investor for a number of rea-
shifts in composition are inevitable at the sons linked to underlying geomorpho-
harvesting intensities required to support graphic controls and the need to maintain
capital-intensive, industrial approaches. competitiveness in a market characterized
by a wide range of interchangeable material
Reduced-impact logging options.

Stand damage control is arguably the raison Sustainability of harvest rates and intensities
d’être for employing reduced-impact log-
ging (RIL) techniques (Hendrison, 1989). A competitive real rate of financial return
However, reducing the unit costs of extrac- on an investment is rarely characterized by
tion is commonly submitted as the financial a minimum 50+ year period of maturity.
pay-off associated with controlling damage Yet, this may be what is required in order to
(Holmes et al., 1999; van der Hout, 1999, sustain timber yields in the Guiana Shield
2000). This pay-off is intimately associated through up-front investments in refined
with the up-front investment in preharvest management practices under operational
Forest Conservation and Management 501

and market conditions at the turn of the threshold of logging intensity. Above this
21st century. Several factors come into play threshold, the relative value of employing
in defining why the term of investment RIL techniques diminishes rapidly as the
should be of such length. density of felling and extraction overrides
First, stand-level increments for many canopy and potential crop-tree conserva-
heavy hardwood timber trees are generally tion effects typically improved by RIL at
some of the lowest in the industry (Prince, lower intensities (Sist et al., 1998; van der
1971; Veillon, 1985; Silva et al., 1995; Hout, 2000). Consequently, a considerable
Zagt, 1997). Even after considerable logging volume of harvestable timber is foregone
intensities, that should have strong liberat- over the short-term as an investment in
ing effects, models indicate that forest longer-term timber quality and sustained
stocking is unlikely to reach preharvest volumes takes precedence. This creates an
levels within 40 years at sites in western almost irreconcilable predicament for large-
tropical Venezuela (Kammesheidt et al., scale operators that work to achieve the
2001), Guyana (Zagt, 1997; ter Steege et al., economies of scale necessary in recouping
2002) and Suriname (de Graaf, 2000). hefty capital investments in equipment and
On this basis, a cutting cycle of at least 60 downstream processing facilities.
years has been recommended in order to Relatively high work rates must be
maintain forest biomass (Zagt, 1997; achieved in order to financially warrant the
Kammesheidt et al., 2001), although the use of heavy machinery and large sawmills.
CELOS system in Suriname originally rec- In Guyana, initial efforts to mechanize the
ommended a 20–25-year cycle (de Graaf, timber industry focused precisely on the
1986). higher work rates that could be achieved
Secondly, hidden defects such as hol- (Grayum, 1971). Expressed on a unit basis,
lowing or decay are very common in com- costs were expected to be lower than more
mercial size-classes, at least in most timber labour-intensive methods (Fig. 9.8A).
trees of Suriname and Guyana (de Milde, However, expressed as a function of time,
1970; de Milde and Inglis, 1974; D. increased use of machinery demanded
Hammond, personal observation). While higher production rates, because daily costs
not normally considered in assessments of were considerably higher (Fig. 9.8B).
forest stand responses to silvicultural inter- In forests across the Guiana Shield,
vention, defect rates among many slower- slow growth rates, high defect rates, size-
growing, heavy hardwood species (e.g. able areas of non-commercial forest (e.g.
Chlorocardium, Mora, Swartzia) are gener- bana, muri scrub, palm swamp) (Fanshawe,
ally greater than in faster-growing, long- 1954; Vink, 1970; Welch, 1975b), poor tim-
lived colonizers that dominate timber ber price growth and increasing transport
industries in many other parts of the distances combine to challenge the finan-
neotropics (e.g. Swietenia, Cedrela, cial sustainability of capital-intensive oper-
Cedrelinga). Defective logs, once mistak- ations. Yet, views persist that much shorter
enly felled, are normally left in the forest or cutting cycles and higher intensities should
at the consolidation market or mill yard be employed across some remote areas of
once detected. Most modern gang-saw oper- the region (J. Leigh, personal communica-
ations do not easily cope with these types of tion).
defects. Thus, the ‘processable’ volume can In reality, commercial logging intensi-
be considerably lower than the standing ties have traditionally been self-limiting
commercial volume often used as a measur- across many forest areas because of these
ing stick to gauge initial extraction opportu- factors. Extraction rates of 6–10 m3/ha are
nity and later postharvest stand not uncommon in many modern, large-
performance. scale operations (van der Hout, 1999;
Thirdly, investment in good forest Armstrong and Inglis, 2000). Yet, even
management practices, such as RIL tech- modest increases to 20 m3/ha recom-
niques, can only yield a return below a mended as part of the CELOS Management
502 D.S. Hammond

Fig. 9.8. The calculated cost of employing chainsaws and mechanized skidders and loaders in forest oper-
ations in central Guyana expressed as a function of volume extracted and operational time. Data source:
Grayum (1971).

System have not precipitated the expected areas. Forests throughout the Guiana Shield
re-stocking necessary for another harvest typically show relatively high dominance
after 20 years, even after liberal application of 1–3 commercial taxa and their distribu-
of refinement treatments (de Graaf et al., tion is always concentrated at scales
1999). defined by edaphic or dispersal limits (see
Low, but highly concentrated, stocking Chapter 7), a feature not as commonly
of timber trees presents specific problems encountered in other commercial forests in
when heavy hardwood regeneration is opti- the neotropics. Thus, the optimal in-
mized under relatively low levels of canopy tensity of logging that makes best use
openness (ter Steege et al., 1995), but opti- of the investment in RIL depends to a great
mal harvest rates require extraction of a extent upon the distribution of commercial
larger number of stems from these smaller stems throughout the stand and how this
Forest Conservation and Management 503

interacts with the balance of financial con- (docks and wharves), and resistance to wear
siderations, such as achievable market (flooring, decking) and maritime infestation
price, amount of rent (taxes, royalties) pay- (locks, docking), among other applications
ments and changing costs of operational (Fig. 9.9). Higher harvesting rates, however,
inputs, such as fuel, parts and staff remu- will inevitably liberate the canopy and
neration. stimulate faster growth. For many forests in
Fourthly, the absence of cost-effective the region, this will increase the rate of
silvicultural techniques also affects the rel- commercial volume accrual, but principally
atively long term of investment required for with faster-growing, light-wooded species.
sustainable timber production in the Heavy hardwood species recruited during
Guiana Shield. Most techniques, including these periods into harvestable size classes
liberation and enrichment planting, have may also see a reduction in wood density as
been shown to yield comparatively better a consequence of growth stimulated by
performance of target timber species, but more open canopy conditions (but see fac-
have also proven far too costly (de Graaf et tors affecting density in ‘Wood density’,
al., 1999). Moreover, decision-making at the Chapter 7 and ‘Growth in relation to canopy
time of postharvest, silvicultural interven- openness’ in Chapter 3). Consequently,
tion may not adequately predict changing future timber production may revolve
market demand and opportunities. around lower-density wood, a market that
Consequently treatment costs may be borne is open to a wide range of species grown
up-front, only to find they were incurred to principally in plantation forests throughout
eliminate species in demand at the time of the world (compare across regions in Fig.
the subsequent harvest. Non-commercial 9.9). Given production levels within the
species may also play an important role in Guiana Shield, this route would not appear
provisioning forest stands with important to provide any comparative advantage over
nutrient import (Perreijn, 2002) or dispersal that already existing through production
services (Hammond et al., 1996) (see of increasingly scarce heavy hardwood
Chapter 4). Vertebrate-dispersal of timber material (Hammond, 1999). The trade-off
tree seeds is arguably greater in forests of required, however, is a focus on lower pro-
the Guiana Shield than in any other tropical duction volumes from forest managed for
region (e.g. Jansen and Zuidema, 2001) and timber through natural regeneration. There
these services form an important part of are, however, other non-timber benefits
natural regeneration approaches to timber attached to forests in the Guiana Shield and
management. The cost-ineffectiveness of these will for the most part continue to
postharvest intervention limits the range of accrue under low-impact systems of timber
tools that can be employed in stimulating production. Compared with much higher
faster growth, if that is the main objective in rates of tropical forest loss elsewhere and
timber management, to the amount of remaining standing forest areas (see
canopy openness created during felling and Table 9.1), continued low-level wood
extraction. production from forests in the Guiana
Shield should see increases in unit prices if
Quality vs. quantity as the timber a focus on wood product quality is main-
management objective tained.

A focus on timber quality would appear Operational waste and inefficiencies


intuitive given the conditions structuring
operational costs and constraints on forest RIL techniques measurably offset financial
productivity across the region (Hammond, investments in planning by reducing opera-
1999). Properties of high structural tional waste and inefficiency. However, for
strength, resistance and durability define the broader forest operation and forestry
most timbers in demand overseas, princi- sector to achieve sustainability (after
pally for their load-bearing capacities Goodland and Daly, 1996), other point
504 D.S. Hammond

16,000
Modulus of rupture
14,000
12,000
mm-1
2
NN/mm

10,000
8000
6000
4000
2000
0
140 Modulus of elasticity
120
100
2 N/mm-1

80
N mm

60
40
20
0
Specific gravity
0.7
0.6
cm-3

0.5
3
gg/cm

0.4
0.3
0.2
0.1
0
W. Africa
C America

SE Asia
Guianas

Western

Temperate
SA

Fig. 9.9. Mean (± SE) wood specific gravity and two of its mechanical property correlates for timber species
grouped by region. See Fig. 7.6 for data sources.

sources of waste and inefficiency must be region in the first place, but these are too
wrangled and subdued if RIL is to achieve complex to be explored here (see Repetto
its measured objective of leaving a growing, and Gillis, 1988). It is clear, however, that
non-depreciated forest without foregoing operational waste and inefficiency in
the immediate financial objective of doing felling and extraction of timber form only
business. Arguably, other factors may be the first links in a chain of waste (Fig. 9.10)
spurring investment in marginal prospects that, depending on its length and strength,
of financial return from timber in the shield can terminate any prospect of positive
Forest Conservation and Management 505

FOREST

Unskidded trees Undetected defects

Skidding damage

Added
volume Bole
needed to sizing
Transport
meet accidents
capital
outlay

MILL
Fungal-insect damage, splitting, drying

Mill conversion Warping,


losses cupping, checking,
bowing, insect &
fungal damage

MARKET

Poor protection,
For Sale theft & accidents,
poorly graded

Fewer sales/lower sale price

Fig. 9.10. Schematic describing the chain of waste that can exist in inefficient and poorly managed timber
production companies. RIL techniques can reduce this waste up to the forest-to-mill phase (arrow). Cost
savings of employing RIL can be rapidly eroded through inflated downstream waste.

financial return to timber operators after to offset initial capital outlay and mounting
meeting reasonable tax burdens.4 As a con- depreciation costs through economies of
sequence, accumulated costs associated with scale. Residual commercial stems in logged-
small, but frequent, wastages downstream over stands in particular can be targeted, par-
can accrue (Fig. 9.11A, B), putting pressure ticularly as transport distance between mill
to sell finished products at lower prices. and unlogged forests increases unit costs and
Ironically, high losses to waste and increas- multiple re-entries, often within the same
ing operational costs can spark further year, become commonplace (D. Hammond,
demand for logs from the forest in an effort personal observation).
506 D.S. Hammond

Fig. 9.11. (A, B) Chasing waste. Extracting timber that is never processed creates unnecessary pressures on
the forest resource since more logs must be extracted in order to meet financial targets. Reducing waste
through transport, storage and milling phases will yield higher returns and improve government revenue.
(C, D) Changing face. The changes to the forest system created by surface mining in the Guiana Shield
can be severe when rehabilitation effort is low. Tall grass in (D) is the common non-flooded savanna
species, Andropogon bicornis.

Discount rates and heavy hardwood growth achieved. Therefore, ecological factors that
and mortality influence the rate of increase should also
shape the rate of asset appreciation, all else
Defining appropriate discount rates to being equal. The primary determinants of
apply when assessing the net present value this rate are based on basic population
(NPV) of commercial forests remains a mat- changes defined by growth and mortality
ter of almost arbitrary assignment. Yet NPV rates. Yet, if growth is slow, as it is for many
is a fundamental calculation influencing heavy hardwoods common to the Guiana
the economic viability of investment in Shield, then the fraction of these lost to
tropical forests. Typically, discount rates mortality events while occupying commer-
are applied as brackets around typical lend- cially harvestable size classes will be con-
ing rates when calculating forest NPV. siderably greater than faster growing
However, the economic factors affecting species with typically shorter life spans (ter
lending rates are hardly sensitive to Steege and Hammond, 1996). This would
changes in the forestry sector, particularly suggest that discount rates based on appre-
across the Guiana Shield where it typically ciation of heavy hardwood timber stocks
accounts for less than 5% of GDP (at factor should be set higher than stands appreciat-
cost). Presumably, when logged forests are ing largely through growth of faster-grow-
viewed (solely) as timber assets, they appre- ing commercial species. Since these
ciate based on the rate of increasing com- invariably form a larger part of future com-
mercial stock, the cost of bringing the mercial growing stocks under natural forest
product to the market and the market price management regimes, discount rates
Forest Conservation and Management 507

applied to forest NPV should decrease with 6. Large-scale surface degradation with lit-
time, all else being equal. Factors that stim- tle or no post-closure restoration effort (Fig.
ulate greater mortality rates, such as cata- 11C,D).
strophic disturbance events, and therefore
greater investment risk and higher lending Mitigation of mining impacts
rates, would suggest spatial variation in the
application of discount rates across geo- Although considerable attention has been
morphographic regions. focused traditionally on large, international
mining consortia that are operating at vari-
ous locations throughout the region (e.g.
Mainly mining Colchester, 1997), small-scale miners are
more numerous, less easily regulated and
The extraction of gold, bauxite, diamonds more often working illegally. When envir-
and iron is arguably the most substantial onmental and social impacts become unac-
non-timber use of forests in the Guiana ceptable, it is virtually impossible to
Shield (see Chapter 8). In terms of biophys- determine accountability in a way that is
ical and socio-economic impact, the pro- more transparent with larger, stationary
duction of minerals spreads across a much operations. Among these impacts, mercury
greater portion of the region, figures more poisoning has arguably attracted the great-
prominently in the agenda of national eco- est attention because of the important pub-
nomic development and conjures up lic health consequences to people living in
greater visions of rapidly accruing wealth rural environments and dependent on
than virtually any other forest-based activ- aquatic resources for their livelihoods. Sing
ity. Most mining operations open and shut et al. (1996) measured significantly higher
over a 15-year period (see ‘Commercial mean organic mercury levels (31.3 mg/l) in
mining’, Chapter 8), or are consumed by blood samples taken from Makuxi villagers
operational inefficiencies as global com- along the Cotingo River with high exposure
modity markets fluctuate independent of to mining operation affluent. Levels of total
extraction costs. The long planning horizon mercury in urine of Maroon gold workers in
of timber production appears almost slug- Suriname were significantly higher than
gish by comparison and mining does not those not involved in gold mining, although
need to carry with it any visage of sustain- blood levels were generally low (de Kom et
ability comparable to the timber industry. al., 1998). In French Guiana, dietary habits
It does, however, increasingly need to accounted for a significant part of mercury-
deflect growing criticism over the way in level variation in hair samples of rural resi-
which it is being managed across the region dents, particularly among Amerindian
and the negative impacts that this is caus- children (Cordier et al., 1998). The dietary
ing through: connection is largely based on fish that
readily sequester and concentrate mercury
1. Little or no management of sediment in their tissues. In Suriname, a survey of the
and mecury effluent; freshwater fish community showed that
2. Rapid spread of a wide variety of com- predatory fish had significantly higher con-
municable diseases, including HIV and centrations of mercury than non-predatory
malaria (Palmer et al., 2002); species and that these levels were often
3. Breakdown of family units structures exceeding maximum permissible concen-
(Heemskerk, 2001); trations (Mol et al., 2001). River bottom sed-
4. Negative impacts on neighbouring rural iments in Guyana tend to concentrate
communities that may have very different mercury downstream from intensive min-
forestland use priorities; ing districts (Miller et al., 2003), where it is
5. Extirpation of local wildlife through available for uptake by benthic feeders.
unrestricted hunting, fishing and live-ani- Although significant background mercury
mal collecting; and levels have been detected throughout many
508 D.S. Hammond

parts of the region as a consequence of its Non-timber forest plant products


long history of weathering and exposure to
forest fires (Veiga et al., 1994; Fostier et al., Historically, a number of important NTFPs
2000; Santos et al., 2001), both important have been commercialized across the
sources of naturally occurring mercury, the Guiana Shield (Table 8.2), including balata
role of mining in catalysing widespread (Manilkara), rosewood oil (Aniba), copal
exposure to much higher concentrations of resin (Hymenaea), sarrapia (cumaru) beans
this toxic heavy metal is difficult to dispute (Dipteryx), curare (Strychnos), among oth-
(Nriagu et al., 1992; Roulet et al., 1999). ers. Depending on the part of the plant
Unfortunately, the hydrochemical proper- required, the harvest of many NTFPs led to
ties of most rivers in the Guiana Shield commercial extinction across much of their
make them ideal transporters of metallo- range in a manner more commonly attrib-
organic complexes of methylated mercury uted to timber production (see Chapter 8).
(Silva-Forsberg et al., 1999). Thus, concepts of sustainable use and
harvest intensity apply equally to both
timber and non-timber products. In the
Mainly non-timber forest products (NTFPs) Guiana Shield, this has become particularly
important in those industries that have
Non-timber approaches to forest use include developed industrial-scale demand for raw
a large basket of potential goods and services. materials.
Many studies do not discriminate commer-
cially viable and subsistence-oriented materi- Palm-heart sustainability
als, particularly when NTFP abundances are
being inventoried. Frequently every plant has Among these, palm-heart (Euterpe spp.)
several useful purposes within the traditional production has received the greatest atten-
use system of most Amerindian people, so tion, both as a growing industry along the
assessing NTFP status alone does not allow Atlantic margin of the Guiana Shield, but
for good discrimination of prospective also with a concern towards the long-term
income-earners. Here an emphasis is placed sustainability of current production meth-
on NTFPs in the Guiana Shield that have or ods. The coastal Atlantic region of the
could generate sufficient market demand to Guiana Shield dominates the geographic
provide some measure of income support to distribution of E. oleracea, the multi-
rural livelihoods. stemmed species that accounts for most
Direct use forms include the produc- palm-heart production from the region. In
tion of plant oils, resins and latexes, fibres contrast, the single-stemmed E. precatoria
for furniture and handicraft manufacture, is the primary source of material for the
food plants, collection of wildlife for the Bolivian palm-heart industry (Bojanic,
pet trade, and fish and bushmeat for market 2001) and prior industries in southern
sale. Biochemical components of plants Brazil centred on the single-stemmed E.
and animals can also be surveyed for their edulis are now largely closed due to com-
prospective use in agro-chemical or phar- mercial extinction of the raw material
maceutical applications. Indirect uses source (Anderson and Jardim, 1989; Pollak
include services provided by forests in the et al., 1995). In fact, sustainable manage-
form of tourism, downstream water quality ment of palm-heart production from single-
and flow regulation, carbon storage and stemmed Euterpe appears largely
sequestration and more complex connec- unachievable at any reasonable timescale
tions associated with regulation of due to slow growth rates (Peña and
land–ocean–atmosphere fluxes forming the Zuidema, 1999) and absence of root sucker-
global biogeochemical cycles. Non-use ing (Anderson, 1988; Johnston, 1995; van
forms include existence value concepts Andel, 1998, 2000). Long-term production
attached to philosophical or spiritual from E. oleracea is feasible if size-based
beliefs supporting forest persistence. harvesting limits, fractional shoot collec-
Forest Conservation and Management 509

tion and adequate fallow periods are Wildlife is the most important com-
adhered to by palm-heart harvesters (Pollak mercial NTFP across the Guianas.
et al., 1995; van Andel, 2000). In most Registered exports of wildlife, composed
cases, a fallow period of 4–5 years appears mainly of psittacines (parrots, macaws),
adequate to ensure stability in the popula- reptiles (snakes, lizards, turtles), monkeys
tion size class distribution. (Cebus, Saguinas, Saimiri) and aquarium
Thus, the industry across the Guiana fish constitute a significant foreign-
Shield has a striking comparative advan- exchange earner and source of income for
tage based on its E. oleracea stands if it can people in Guyana, Suriname and the Rio
be managed sustainably. Competition in Negro region of Brazil (Chao and Prang,
Euterpe heart production is restricted to 1997; De Souza, 1997; Duplaix, 2001;
areas in and around the Amazon estuary Ouboter, 2001).
where palm-heart production must also Despite its prominent figuring in NTFP
compete with other profitable uses of revenue generation across the region,
Euterpe that do not currently generate any wildlife use remains one of the least regu-
significant domestic demand across the lated of forest-based activities and little
Guianas. Competition from palm-heart pro- information is available regarding popula-
duction in Bactris gasipaes, a semi-domes- tion sizes, rates of use or how long these
ticate found throughout the neotropics can be sustained. All indications suggest
(Henderson et al., 1995) is more substantial. that use is more targeted, less widespread
According to Arkcoll and Clement (1989, and less intensive compared to other low-
cited in van Andel, 2000), this palm is land neotropical regions that have been vir-
capable of producing palm hearts at six tually defaunated, but productivity in the
times the rate of E. oleracea. shield is also typically much lower and lim-
Palm-heart production appears capable its defined elsewhere are unlikely to be
of bring a modicum of livelihood support to applicable across geomorphographic
rural communities across the northern regions. Some species have already been
shield perimeter if strict, but simple, har- clearly over-harvested historically, since
vesting rules are followed and enforced. It their geographic ranges have contracted to
also provides a source of income that is close more remote locations (e.g. black caiman,
to their home communities, avoiding many arapaima). Sociocultural conditions are
of the social impacts associated with men also different. For example, monkeys are
leaving to work for extended periods away rarely hunted for bushmeat in Guyana,
from home, a feature commonly associated although consumptive rates across other
with small-scale, gold-mining operations. parts of the neotropics are often as high or
higher for primates than other taxa
Wildlife use (Redford, 1993).
Unregulated collection of wildlife to
The socio-economic issues attached to meet a growing commercial trade in bush-
wildlife use in tropical forests can be excep- meat, fish and live specimens is arguably
tionally complex, as a wide variety of per- the most important challenge facing verte-
spectives and objectives view forest brate conservation and sustainable use
animals in very different ways (Freese, across the Guiana Shield. Significant con-
1998). Forest wildlife use is seen through a cerns remain regarding how best to estab-
wide range of perspectives ranging from its lish functional quotas on the wildlife trade
value as a protein source, as the focus of an (Pilgrim, 1993) and the impact this trade
esoteric, but potentially important, culture has on rural societies (Forte, 1989), but the
of exotic pet-keeping through to their spiri- largest issue confronting the future of
tual representations. The catalysts and con- wildlife across the region is the virtual
sequent patterns of commercial wildlife use absence of any restrictions on hunting and
are, however, surprisingly similar (Fimbel fishing season, quantity or location
et al., 2001). (Duplaix, 2001). Unmitigated access to
510 D.S. Hammond

(mainly) public wildlife resources without lation that remains unpaved is the
any payment of taxes or rents similar to Oiapoque to Macapa road in Amapá and
those placed on plant products invariably this is earmarked for paving under the
stimulates wasteful commercial consump- Avança Brasil programme (Carvalho et al.,
tion of wildlife because it creates opportu- 2001).
nities to make a solid return on investment Pavement of the Georgetown to Lethem
without the need to manage operations crit- road would connect all of the major cities
ically. Like logging, a chain of waste ensues along the shield periphery, Manaus, Ciudad
from point of harvest to final market. Bolivar, Georgetown, Paramaribo and
Income derived from the use of wildlife that Cayenne, in a way that would accommo-
should bolster public revenues is instead date much larger volumes of traffic carrying
taken as super-profit by operators or lost as much larger payloads. The feasibility of
wastage. Ultimately, local and national such an undertaking has been explored
communities bear the cost of over-exploita- (Environmental Resources Management,
tion through a decreasing subsistence share 1995), although the advantage of upgrading
and lost government revenues. At the the 585 km length of road to the resident
national scale, the absence of any restric- Guyanese population of less than 750,000
tions on hunting and fishing is difficult to remains largely unclear in comparison to
resolve with interest in developing viable the relative ease of access afforded Brazil’s
ecotourism industries, maintaining an effi- northernmost state, Roraima, with its popu-
cient, profitable and legal export of live lation of 325,000 (Forte, 1990; Forte and
specimens and protecting local rural com- Benjamin, 1993).
munity subsistence use needs. All of these Roads across remote frontiers act as
represent real comparative, but largely conduits for the legal transport of people
unmarketed, advantages held in the region and goods that can broaden cash income-
due to the high degree of forest integrity earning opportunities for rural households
that has been maintained relative to other through greater market access for their agri-
regions that were capable of providing sim- cultural and timber products. But roads
ilar goods and services. also can act as very effective thoroughfares
for the transmission of disease, invasive
and exotic pests, smuggling, poaching of
Managing forest roads and road-building wildlife, timber and other plant materials,
unpermitted mining, violence and crime.
If lack of regulations, restrictions and their Massive influxes of cash and working capi-
effective enforcement can critically impede tal can rapidly overwhelm and then erode
progress towards sustainable forest manage- existing rural community social structures.
ment across the Guiana Shield, then an For example, even modest increases in traf-
expanding road network is the main agent fic along the Georgetown to Lethem road
capable of ensuring that it fails altogether. running through the Iwokrama Forest saw
This is not to engender the view that signif- an increase in illegal hunting activities,
icant benefit cannot be derived from road timber poaching and conflicts regarding
establishment, but to emphasize that in the uses of roadside creeks that supplied water
absence of effective road management, to adjoining communities (Ousman, 1999).
costs attached to their use will invariably More substantial impacts are well-estab-
outweigh the benefits roads can deliver at lished for roads built in other parts of the
local, national and regional scales. The Amazon, largely due to an absence of any
main cities of the region were connected by road management presence capable of
all-year roads by 2002 with the upgrading organizing where and how roadside land is
of the Georgetown to Lethem road in used and mitigating impacts created by
Guyana, although this remains unpaved road-users. The Guiana Shield remains one
(Fig. 9.12). The only other road prospec- of the last forest frontiers largely due to its
tively connecting the main centres of popu- inaccessibility. Improving access will bring
Forest Conservation and Management 511

Fig. 9.12. Major road arteries connecting large towns and cities across the Guiana Shield and beyond.
Note the pivotal position of Boa Vista in the overland flow of traffic between the Guianas, Brazil and
Venezuela. Road development between Guyana and Venezuela (through Cuyuni-Nuria) (filled arrow) and
French Guiana and Amapá (empty arrow) would diversify the transit access across the region.

further opportunities for economic devel- underlying environmental uniformity and


opment only if these roads are effectively perpetuated an implicit belief that all tropi-
managed to prevent or mitigate the wide cal forests are the product of similar past
range of negative impacts that can follow histories and uniformly respond (nega-
road development. tively) to the same basket of destructive
land-use practices. Selective extraction and
forest clearance approaches to timber pro-
Management Scale, Focus and Objective duction get muddled into a single mass.
– a Perspective Non-wood products equally begin to form a
conglomerate of economic alternatives.
Scale Shifting agriculturalists are melded into a
single group that ignores important varia-
Environmental concerns over the use of tion in the marginal utility they receive,
tropical forests in many ways have been impacts they exact and core competencies
transfixed at the largest inclusive scales, but through which they manage forest
consistently studied at some of the smallest resources. The scale of mining operations
scales. In fact, scaling-up interpretation of and the realistic costs and benefits attached
many site- and time-specific results has to these different levels are not always ade-
unintentionally cemented the notion of quately considered. Thus studying the
512 D.S. Hammond

socio-economic and environmental conse- the political will is sufficient to enact legis-
quences of commercial-scale use of in situ lation mandating appropriate forms of sus-
forestlands from a presumptive perspective tainable management and allow this
of either net benefit or costs without faith- legislation to be actively implemented.
fully assessing the other can further foster Historically, where large companies have
irrational decision-making and add a dominated commodity economies, the
polemical coating to problem-solving excessive weight of their political leverage
efforts. In the end, these efforts should seek has led to large-scale devaluation of forest-
to bring the maximum net benefit to the lands and minimized the flow of benefits as
widest cross-section of society. Implicit a result of their activities. A focus on either
scaling-up of single case history examples small- or large-scale routes to forest use can
of success and failure or research findings is lead to similar outcomes, only through dif-
often inappropriate when variation in bio- ferent routes to management failure.
physical and socio-economic features at
these larger scales is considered. Getting
the scale width of applicability correct is How does management fail?
the first step in identifying the necessary
tolerance (or robustness) of forest manage- Few aspirations, few opportunities
ment focus in the Guiana Shield.
The focus of forestland use is also shaped
by human aspirations and these, in turn, by
Focus prevailing social norms at the time of deci-
sion-making. Education, health, communi-
Management focus must be sufficiently cation, transport and the technologies that
robust to weather the vagaries of larger- affect changes in living conditions can alter
scale socio-economic and biophysical fluc- these social norms and aspirations attached
tuations. Chronic boom-and-bust cycling of to forest use. When opportunities are few,
commodity production that has typified aspirations are limited and concepts of
economies of the Guiana Shield creates long-term management for future forest
enormous employment and inflationary rip- value rarely figure in the extremely short-
ples that deflect progress towards social term planning horizons of most people liv-
and economic stability. ing in the commodity-driven economies of
The resulting instabilities can drive the Guiana Shield.
further environmental degradation. Bust Extending the planning horizon is an
phases are dominated by large numbers of important precondition to sustainable man-
small operators seeking to meet their short- agement, but this can only occur where eco-
term subsistence needs with very little nomic and political processes culture a
capacity or motivation to minimize collat- sense within both public and private sec-
eral damage (e.g. Heemskerk, 2001). Their tors that forests deliver a wider and more
activities are difficult to regulate and man- permanent range of benefits. The distribu-
age. The sum total of their individual tion of benefit as a result of forest use also
impacts creates a ‘tragedy of the commons’ influences perspectives on acceptable lev-
scenario where compounded use of the els of waste and extraction intensity. Those
same resources leads to rapid commercial deriving the least direct benefit and bearing
extinction and options for sustainable man- the greatest indirect costs from extraction
agement are limited to restoration. are most likely to be the most conservative
Alternatively, large-scale industrial in their perspectives of acceptable levels.
extraction of resources by a few corporate This implies that the focus can change with
operators can ease and reduce the cost of level of ownership, rather than necessarily
monitoring for sustainable management if with the scale of management.
Forest Conservation and Management 513

Mismatched land-use approaches Objectives

Equally, the mismatch of land-use If ownership influences acceptability of for-


approach with biophysical features further est use and low-energy systems govern for-
drives gross leakages that slacken the iner- est ecosystem change, then the objective of
tia needed to achieve socio-economic management at the largest scale should be
development in commodity-driven an inclusive system that offers structured
economies. Large-scale forest clearance for opportunities for a wide range of
agriculture or livestock on sand-dominated approaches tailored to different prospective
tropical soil facies leaves few livelihood types of ownership. At smaller scales,
alternatives. Once these are cleared and working approaches should integrate core
abandoned, an extremely slow forest re- conservation measures to mitigate energy
establishment process modulated by nutri- imports and identify optimal harvesting
ent evacuation, large-seed dispersal intensities that will meet reasonable, short-
distances and inherently slow growth and medium-term financial objectives with-
capacities (see Chapter 3) renders these out liquidating longer-term opportunities
areas of marginal economic potential. through accelerated forest decline.
These sand-dominated soils occupy large Establishing and enforcing allowable
areas of the region (and define several of the boundaries to the way in which the
endemic ecoregional units) (see resource can be managed constitutes the
‘Ecoregions’, this chapter, and ‘Soils’, main approach to achieving this objective.
Chapter 2). Public forest-use policy and the agencies
In the Guiana Shield, biophysical tra- responsible for its implementation consti-
jectories are defined by low-energy systems tute the main tools for ensuring that long-
(see Chapter 7). If the focus of forest land term opportunities are not unnecessarily
use is sustainable delivery of specific bene- liquidated and that the forest resource
fits without erosion of future opportunities, industry is not beset with unrealistic or
then adopted systems must also seek con- uneconomic conservation measures.
servation measures that reduce their role as Maintaining this balance in the low-energy
external importers of energy into the environment of the Guiana Shield repre-
ecosystem. Across most land-use options, sents one of the greatest challenges to sus-
these conservation measures are a function tainable management of tropical forests
of extraction intensity, operational ineffi- anywhere. Sustainable use and conserva-
ciencies or both. Whether seen as collateral tion of forests in the region will in the end
stand damage, non-selective wildlife har- depend on the international and national
vesting, uncontrolled anthropogenic fire, support for the work of these agencies and
poor road-building or unmitigated mining the commitment of their staff in making
effluent, the injection of energy into the sys- best use of this support in achieving objec-
tem accelerates the slide down an intrinsic tives that benefit wider society. How effi-
dampening slope of forest life-expectancy. ciently natural capital is transformed into
At a geological timescale, forests of the social capital through directed public and
Guiana Shield are already in a state of tran- private sector investment will determine
sitory decline and the application of man- the role that regional forests will play in
agement techniques adopted from other national development or, sadly, chronic
regions that are currently in a state of tran- under-development. How people in the
sitory ascent (e.g. western Amazonia, cen- Guiana Shield ultimately define a lost
tral America, SE Asia), rather than descent, opportunity will and should determine
will invariably accelerate this decline. how long and how much utility they and
their future generations can expect to enjoy
from their forests.
514 D.S. Hammond

Notes
1 This area incorporates the Jaú National Park (see Fig. 9.1 for location), Mamirauá and Amana
Sustainable Development Reserves and the Anavilhanas Ecological Station (WCPA – World
Database on Protected Areas).
2 How the latter criterion was spatially discriminated is not clear to this author from the avail-

able literature.
3 Explanatory variables used in the analysis appear to violate assumption of orthogonality and

are subject to unknown error structure (Vanclay, 2001). One could just as easily argue that
enforcement is proven most effective in maintaining tropical park integrity because other efforts
to adequately pair protection with expanding livelihood opportunities and effort to demonstrate
to people the purpose and benefits of habitat protection have been neglected or failed. In fact,
degradation of adjoining habitat and effectiveness of guard density are not independent and may
be strongly correlated. A more appropriate conclusion would have stated that individuals
involved in the establishment, management and funding of tropical parks believe that guard den-
sity is the best predictor of park effectiveness. If local community residents did not form part of
the survey, it is difficult to see how local support and participation variables were calculated
independent of opinions expressed by park-associated respondents who may have quite differ-
ent views on park performance and attribution of causality.
4 This normal tax burden effectively equates to the appropriate level of economic rent sought by

the owner. In the case of the Guiana Shield, where most commercial forests are owned by the
public and stewarded on their behalf by the relevant government agency, this rent would be the
difference between revenues and the sum of the production costs and a normal profit (often
around 30%). In effect, it is a rent for use of the land in its mature, timber-bearing condition.
Often rent capture has been well below this difference and is believed to have formed one of
the more important incentives for otherwise inefficient logging practices to persist. Rents,
through the various forms they can take, have traditionally been very low in most countries in
the Guiana Shield (e.g. Oliphant, 1938; Palmer, 1996; Whiteman, 1999).

References

Agriconsulting (1993) Preparatory Study for the Creation of a Protected Area in the Kanuku Mountains
Region of Guyana. Agriconsulting, Rome.
Anderson, A.B. (1988) Use and management of native forests dominated by açai palm (Euterpe oleracea
Mart.) in the Amazon estuary. Advances in Economic Botany 6, 144–154.
Anderson, A.B. and Jardim, M.A.G. (1989) Costs and benefits of floodplain forest management by rural
inhabitants in the Amazon estuary: a case study of Açai palm production. In: Browder, J.O. (ed.) Fragile
Lands of Latin America: Strategies for Sustainable Development. Westview Press, London, pp.
114–129.
Arkcoll, D. B. and Clement, C. (1989) Potential new crops for the Amazon. In: Wickens, G.E., Haq, N. and
Day, P. (eds) New Crops for Food and Industry. Chapman and Hall, London, pp. 150–165.
Armstrong, S. and Inglis, C.J. (2000) RIL for real: introducing reduced impact logging techniques into a com-
mercial forestry operation in Guyana. International Forestry Review 2, 17–23.
Barbier, E.B., Burgess, J.C., Bishop, J. and Aylward, B. (1994) The Economics of the Tropical Timber Trade.
Earthscan Publications, London.
Barreto, P., Amaral, P., Vidal, E. and Uhl, C. (1988) Costs and benefits of forest management for timber pro-
duction in eastern Amazonia. Forest Ecology and Management 108, 9–26.
Bojanic, A.J. (2001) Balance is Beautiful: Assessing Sustainable Development in the Rain Forests of the
Bolivian Amazon. PROMAB Scientific Series 4. PROMAB, Utrecht, The Netherlands.
Boltz, F., Holmes, T.P. and Carter, D.R. (2003) Economic and environmental impacts of conventional and
reduced-impact logging in tropical South America: a comparative review. Forest Policy and Economics
5, 69–81.
Forest Conservation and Management 515

British Guiana Forest Department (1940–1965) Annual Report of the Forest Department. Department of
Lands and Mines, Georgetown, Guyana
Bruner, A.G., Gullison, R.E., Rice, R.E. and da Fonseca, G.A.B. (2001) Effectiveness of parks in protecting
tropical biodiversity. Science 291, 125–128.
Bryant, D., Nielsen, D. and Tangley, L. (1997) The Last Frontier Forests – Ecosystems and Economies on the
Edge. World Resources Institute, Washington, DC.
Carvalho, G., Barros, A.C., Moutinho, P. and Nepstad, D. (2001) Sensitive development could protect
Amazonia instead of destroying it. Nature 409, 131.
Chao, N.L. and Prang, G. (1997) Project Piaba – towards a sustainable ornamental fishery in the Amazon.
Aquarium Sciences and Conservation 1, 105–111.
Chape, S., Blyth, S., Fish, L., Fox, P. and Spalding, M. (2003) 2003 United Nations List of Protected Areas.
IUCN/UNEP-WCMC, Cambridge, UK.
Colchester, M. (1997) Guyana: Fragile Frontier. Earthscan Publications, London.
Colinvaux, P.A., De Oliveira, P.E., Moreno, J.E., Miller, M.C. and Bush, M.B. (1996) A long pollen record
from lowland Amazonia: forest and cooling in glacial times. Science 274, 85–88.
Coomes, D.A. and Grubb, P.J. (1996) Amazonian caatinga and related communities at La Esmeralda,
Venezuela: forest structure, physiognamy and floristics, and control by soil factors. Vegetatio 122,
167–191.
Cooper, A. (1979) Muri and white sand savannah in Guyana, Surinam and French Guiana. In: Specht, R.L.
(ed.) Heathlands and Related Shrublands. Elsevier, Amsterdam, pp. 471–481.
Cordier, S., Grasmick, C., Paquier-Passelaigue, M., Mandereau, L., Weber, J.P. and Jouan, M. (1998) Mercury
exposure in French Guiana: levels and determinants. Archives of Environmental Health 53, 299–303.
Davis, S.D., Heywood, V.H., Herrera-McBryde, O., Villa-Lobos, J. and Hamilton, A.C. (1997) Centres of
Plant Diversity. A Guide and Strategy for their Conservation. Volume 3 – the Americas. WWF-IUCN,
Gland, Switzerland.
Davis, T.A.W. and Richards, P.W. (1933) The vegetation of Moraballi creek, British Guiana: an ecological
study of a limited area of tropical rain forest. I. Journal of Ecology 21, 350–384.
de Graaf, N.R. (1986) A Silvicultural System for Natural Regeneration of Tropical Rain Forest in Suriname.
Pudoc/WAU, Wageningen, The Netherlands.
de Graaf, N.R. (2000) Reduced impact logging as part of the domestication of neotropical rainforest.
International Forestry Review 2, 40–44.
de Graaf, N.R., Poels, R.L.H. and van Rompaey, R.S.A.R. (1999) Effect of silvicultural treatment on growth
and mortality of rainforest in Surinam, over long periods. Forest Ecology and Management 124,
123–135.
de Kom, J.F., van der Voet, G.B. and de Wolff, F.A. (1998) Mercury exposure of maroon workers in the small
scale gold mining in Suriname. Environmental Research 77, 91–97.
de Milde, R. (1970) Assessment of Hidden Defects on Standing Trees During Inventory Work. Technical
Report No. 4 (SF/GUY 9). FAO/UNDP, Georgetown, Guyana.
de Milde, R. and Inglis, C.J. (1974) Defect Assessment of Standing Trees. FAO Project Working Document
No. 14 FO/SF/SUR/71/506. FAO, Paramaribo, Suriname.
De Souza, B. (1997) Towards a Methodology for the Economic Valuation of Non-timber Forest Products: A
View from Guyana. GNRA, Georgetown, Guyana.
Denevan, W.D. (1970) Aboriginal drained field cultivation in the Americas. Science 169, 647–654.
Dickinson, M.B., Dickinson, J.C. and Putz, F.E. (1996) Natural forest management as a conservation tool in
the tropics: divergent views on possibilities and alternatives. Commonwealth Forestry Review 75,
309–315.
Duivenvoorden, J.F. and Lips, J.M. (1995) A Land–Ecological Study of Soils, Vegetation and Plant Diversity
in Colombian Amazonia. Tropenbos Foundation, Wageningen, The Netherlands.
Duplaix, N. (2001) Evaluation of the Animal and Plant Trade in the Guianas – Preliminary Findings. WWF-
Guianas, Cayenne, French Guiana.
Dykstra, D. and Heinrich, R. (eds) (1996) Forest Codes of Practice – Contributing to Environmentally Sound
Forest Operations. FAO Forestry Paper 133. FAO, Rome.
Environmental Resources Management (1995) Environmental and Social Impact Assessment: Linden-Lethem
Road, Guyana. ERM, London.
Evans, C. and Meggers, B.J. (1960) Archeological Investigations in British Guiana. Bulletin of the Bureau of
American Ethnology, 177. Smithsonian Institution Press, Washington, DC.
FAO (2001) Global Forest Resources Assessment 2000. FAO Forestry Paper 140. FAO, Rome.
516 D.S. Hammond

Fanshawe, D.B. (1952) The Vegetation of British Guiana: A Preliminary Review. Institute paper 29. Imperial
Forestry Institute, Oxford, UK.
Fanshawe, D.B. (1954) Forest types of British Guiana. Caribbean Forester 15, 73–111.
Fimbel, R.A., Grajal, A. and Robinson, J.G. (eds) (2001) The Cutting Edge – Conserving Wildlife in Logged
Tropical Forests. Biology and Resource Management Series. Columbia University Press, New York.
Forte, J. (1989) Wildlife, timber and balata. In: Forte, J. (ed.) The Material Culture of the Wapishana People
of the South Rupununi Savannahs in 1989. ARU – University of Guyana, Turkeyen, pp. 41–44.
Forte, J. (1990) A Preliminary Assessment of the Road to Brazil. ARU – University of Guyana. Turekeyen.
Forte, J. and Benjamin, A. (1993) The Road from Roraima State. ARU – University of Guyana, Turkeyen.
Fostier, A.H., Forti, M.C., Guimaraes, J.R.D., Melfi, A.J., Boulet, R., Espirito Santo, C.M. and Krug, F.J. (2000)
Mercury fluxes in a natural forested Amazonian catchment (Serra do Navio, Amapa state, Brazil).
Science of the Total Environment 260, 201–211.
Freese, C.H. (1998) Wild Species as Commodities. Island Press, Washington, DC.
Funk, V.A. and Richardson, K.S. (2002) Systematic data in biodiversity studies: use it or lose it. Systematic
Biology 51, 303–316.
Funk, V.A., Fernanda Zermoglio, M. and Nazir, N. (1999) Testing the use of specimen collection data and
GIS in biodiversity exploration and conservation decision-making in Guyana. Biodiversity and
Conservation 8, 727–751.
Furch, K. (1984) Water chemistry of the Amazon basin: the distribution of chemical elements among fresh-
waters. In: Sioli, H. (ed.) The Amazon. Limnology and Landscape Ecology of a Mighty Tropical River
and its Basin. Dr. W. Junk, Dordrecht, The Netherlands, pp. 167–195.
Gibbs, A.K. and Barron, C.N. (1993) The Geology of the Guiana Shield. Oxford University Press, Oxford, UK.
Goodland, R. and Daly, H. (1996) Environmental sustainability: universal and non-negotiable. Ecological
Applications 6, 1002–1017.
Goulding, M., Carvalho, M.L. and Ferreira, E.G. (1988) Rio Negro: Rich Life in Poor Water. SPB Academic,
The Hague, The Netherlands.
Grayum, G. (1971) Logging and Forest Management. UNDP Technical Report 12 (Guyana). UNDP,
Georgetown, Guyana.
Hammond, D.S. (1999) Quality vs. Quantity in the Quest for Sustainable Forest Management in the Guiana
Shield. Sustainable Forest-Based Business Partnerships, Georgetown, Guyana, Iwokrama International
Centre.
Hammond, D.S. and Brown, V.K. (1995) Seed size of woody plants in relation to disturbance, dispersal, soil
type in wet neotropical forests. Ecology 76, 2544–2561.
Hammond, D.S. and ter Steege, H. (1998) Propensity for fire in the Guianan rainforests. Conservation Biology
12, 944–947.
Hammond, D.S., Gourlet-Fleury, S., van der Hout, P., ter Steege, H. and Brown, V.K. (1996) A compilation
of known Guianan timber trees and the significance of their dispersal mode, seed size and taxonomic
affinity to tropical rain forest management. Forest Ecology and Management 83, 99–106.
Hammond, D.S., van der Hout, P., Zagt, R., Marshall, G., Evans, J. and Cassells, D.S. (2000) Benefits, bot-
tlenecks and uncertainties in the pantropical implementation of reduced impact logging techniques.
International Forestry Review 2, 45–53.
Heemskerk, M. (2001) Do international commodity prices drive natural resource booms? An empirical analy-
sis of small-scale gold mining in Suriname. Ecological Economics 39, 295–308.
Henderson, A., Galeano, G. and Bernal, R. (1995) Field Guide to the Palms of the Americas. Princeton
University Press, Princeton, New Jersey.
Hendrison, J. (1989) Damage-controlled Logging in Managed Tropical Rain Forest in Suriname. WAU,
Wageningen, The Netherlands.
Heyligers, P.C. (1963) Vegetation and soil of a white-sand savanna in Suriname. Verhandelingen der
Koninklijke Nederlandse Akademie van Wetenschappen, Afd. Natuurkunde 54, 1–148.
Hohenkerk, L.S. (1922) A review of the timber industry of British Guiana. Journal of the Board of Agriculture
of British Guiana 15, 2–22.
Holmes, T.P., Blate, G.M., Zweede, J.C., Pereira, R.J., Barreto, P., Boltz, F. and Bauch, R. (1999) Financial
Costs and Benefits of Reduced-impact Logging Relative to Conventional Logging in the Eastern Amazon.
Phase 1 final report. TFF/USFS, Belem.
Hoosein, M. (1996) A Review of Protection Criteria in Previous Documentation for Protected Areas in
Guyana. Guyana Environmental Protection Agency, Turkeyen, Guyana.
Hubbell, S.P., Foster, R.B., O’Brien, S.T., Harms, K.E., Condit, R., Wechsler, B., Wright, S.J. and de Lao, S.L.
Forest Conservation and Management 517

(1999) Light-gap disturbances, recruitment limitation, and tree diversity in a neotropical forest. Science
283, 554–557.
Huber, O. (1995) Conservation of the Venezuelan Guayana. In: Berry, P.E., Holst, B.K. and Yatskievych, K.
(eds) Flora of the Venezuelan Guayana. Timber Press, Portland, Oregon, pp. 193–218.
Huber, O. (1997) Pantepui region of Venezuela. In: Davis, S.L.D., Heywood, V.H., Herrera-MacBryde, O.,
Villa-Lobos, J. and Hamilton, A.C. (eds) Centres of Plant Diversity – A Guide and Strategy for their
Conservation. Vol. 3. The Americas. World Wildlife Fund for Nature (WWF) and The World
Conservation Union-IUCN. IUCN Publications Unit, Cambridge, UK, pp. 308–311.
Huber, O. (2001) Conservation and environmental concerns in the Venezuelan Amazon. Biodiversity and
Conservation 10, 1627–1643.
IBGE (2000) Cidades@. IBGE (http://www.ibge.gov.br/cidadesat).
IBGE (2002) Banco de dados agregados-extração vegetal. IBGE (http://www.sidra.ibge.gov.br).
Im Thurn, E.F. (1883) Among the Indians of Guiana. Kegan Paul, Trench and Co., London.
Jansen, P.A. and Zuidema, P. (2001) Logging, seed dispersal by vertebrates, and natural regeneration of trop-
ical timber trees. In: Fimbel, R., Grajal, A. and Robinson, J.G. (eds) The Cutting Edge – Conserving
Wildlife in Logged Tropical Forests. Columbia University Press, New York, pp. 35–59.
Johnston, D.V. (1995) Report on the palm cabbage industry in northwest Guyana. In: Forte, J. (ed.)
Indigenous Use of the Forest – Situation Analysis with Emphasis on Region 1. ARU – University of
Guyana, Georgetown, pp. 62–67.
Jonkers, W.B.J. and Schmidt, P. (1984) Ecology and timber production in tropical rain forest in Suriname.
Interciencia 9, 290–298.
Kalliola, R.J., Jokinen, P. and Tuukki, E. (1999) Fluvial dynamics and sustainable development in upper Rio
Amazonas, Peru. In: Padoch, C., Ayres, J.M., Pinedo-Vasquez, M. and Henderson, A. (eds) Várzea –
Diversity, Development and Conservation of Amazonia’s Whitewater Floodplains. NYBG Press, New
York, pp. 271–282.
Kammesheidt, L., Köhler, P. and Huth, A. (2001) Sustainable timber harvesting in Venezuela: a modelling
approach. Journal of Applied Ecology 38, 756–770.
Kennedy, L.M. and Horn, S.P. (1997) Prehistoric maize cultivation at the La Selva biological station, Costa
Rica. Biotropica 29, 368–370.
King, K.F.S. (1963) The role of the forest in the future economy of British Guiana. Commonwealth Forestry
Review 42, 237–241.
King, K F.S. (1965) The use of arboricides in the management of tropical high forest. Turrialba 15, 35–39.
Ledru, M.-P., Blanc, P., Charles-Dominique, P., Fournier, M., Martin, L., Riera, B. and Tardy, C. (1997)
Reconstituion palynologique de la forêt guyanaise au cours des 3000 dernières années. Comptes Rendu
de l’Academie des Sciences, Paris 324, 469–476.
Lescure, J.-P. and Boulet, R. (1985) Relationships between soil and vegetation in a tropical rain forest in
French Guiana. Biotropica 17, 155–164.
Liu, K.-b. and Colinvaux, P.A. (1985) Forest changes in the Amazon basin during the last glacial maximum.
Nature 318, 556–557.
Lugo, A.E. (1995) Management of tropical biodiversity. Ecological Applications 5, 956–961.
Mackay, M.S. (1926) The durability of Greenheart. Journal of the Board of Agriculture of British Guiana 19(1),
10–12.
McNabb, K.L. and Wadouski, L.H. (1999) Multiple rotation yields for intensively managed plantations in the
Amazon basin. New Forests 18, 5–15.
McTurk, M. (1882) Notes on the forests of British Guiana. Timehri 1, 173–215.
Meggers, B.J. (1996) Amazonia – Man and Culture in a Counterfeit Paradise. Revised edition. Smithsonian
Institution Press, London.
Miller, J.R., Lechler, P.J. and Bridge, G. (2003) Mercury contamination of alluvial sediments within the
Essequibo and Mazaruni river basins, Guyana. Water, Air and Soil Pollution 148, 139–166.
Mittermeier, R.A., Mittermeier, C.G., Brooks, T.M., Pilgrim, J.D., Konstant, W.R., da Fonseca, G.A.B. and
Kormos, C. (2001) Wilderness and biodiversity conservation. Proceedings of the National Academy of
Sciences USA 100, 10309–10313.
Mol, J.H., Ramlal, J.S., Lietar, C. and Verloo, M. (2001) Mercury contamination in freshwater, estuarine, and
marine fishes in relation to small-scale gold mining in Suriname, South America. Environmental
Research 86, 183–197.
Moran, E.F. (1995) Rich and poor ecosystems of Amazonia: an approach to management. In: Nishizawa, T. and
Uitto, J.I. (eds) The Fragile Tropics of Latin America. United Nations University Press, Tokyo, pp. 45–67.
518 D.S. Hammond

Myers, N., Mittermeier, R.A., Mittermeier, C.G., Fonseca, G.A.B. and Kent, J. (2000) Biodiversity hotspots for
conservation priorities. Nature 403, 853–858.
Nriagu, J.O., Pfeiffer, W.C., Malm, O., Souza, C.M.M. and Mierle, G. (1992) Mercury pollution in Brazil.
Nature 356, 389.
OCEI/MARNR (2000) Densidad demográfica. OCEI (http://www.ocei.gov.ve).
Oliphant, F.M. (1938) The Commercial Possibilities and Development of the Forests of British Guiana. The
Argosy Co. Ltd, Georgetown, Guyana.
Olson, D.M., Dinerstein, E., Wikramanayake, E.D., Burgess, N.D., Powell, G.V.N., Underwood, E.C.,
D’Amico, J.A., Itoua, I., Strand, H.E., Morrison, J.C., Loucks, C.J., Allnutt, T.F., Ricketts, T.H., Kura, Y.,
Lamoreux, J.F., Wettengel, W.W., Hedao, P. and Kassem, K.R. (2001) Terrestrial ecoregions of the
world: a new map of life on Earth. Bioscience 51, 933–938.
Ouboter, P. (2001) Assessment of the Traded Wildlife Species. WWF-Guianas, Paramaribo, Suriname.
Ousman, S. (1999) A socio-economic and biophysical assessment of the Iwokrama road corridor, Guyana.
MSc thesis, University of West Indies – Mona, Kingston.
Palmer, C.J., Validum, L., Loeffke, B., Laubach, H.E., Mitchell, C., Cummings, R. and Cuadrado, R.R. (2002)
HIV prevalence in a gold mining camp in the Amazon region, Guyana. Emerging Infectious Diseases 8,
330–331.
Palmer, J.R. (1986) Lessons for land managers in the tropics. Revue Bois et Forêts des Tropiques 212, 16–27.
Palmer, J.R. (1996) Report on Forest Revenue Systems. GFC DfID Support Project, Oxford, UK.
Pearce, D., Putz, F.E. and Vanclay, J.K. (2003) Sustainable forestry in the tropics: panacea or folly? Forest
Ecology and Management 172, 229–247.
Peña, M. and Zuidema, P. (1999) Limitaciones demográficas para el aprovechamiento sostenible de Euterpe
precatoria para producción de palmito en dos tipos de bosque de Bolivia. Ecología en Bolivia 33, 3–21.
Perreijn, K. (2002) Symbiotic Nitrogen Fixation by Leguminous Trees in Tropical Rain Forest in Guyana.
Tropenbos-Guyana Series 11. Tropenbos-Guyana, Georgetown.
Pilgrim, K. (1993) Guyana’s wildlife trade. Guyana Review 5, 8–10.
Piperno, D. (1990) Fitolitos, arqueología y cambios prehistóricos de la vegetación en un lote de cincuenta
hectáreas de la isla de Barro Colorado. In: Leigh, E.G.J., Rand, A.S. and Windsor, D.M. (eds) Ecología
de un Bosque Tropical. STRI, Balboa, Panama, pp. 153–156.
Plumptre, A. (1995) The importance of ‘seed trees’ for the natural regeneration of selectively logged tropical
forest. Commonwealth Forestry Review 74, 253–258.
Pollak, H., Mattos, M. and Uhl, C. (1995) A profile of palm heart extraction in the Amazon estuary. Human
Ecology 23, 357–386.
Prince, A.J. (1971) The rate of growth of Greenheart (Ocotea rodiaei Schomb.). Commonwealth Forestry
Review 52, 143–146.
Putz, F.E. (1991) Silvicultural effects of lianas. In: Putz, F.E. and Mooney, H.A. (eds) The Biology of Vines.
Cambridge University Press, Cambridge, UK, pp. 493–501.
Pyne, S.J. (2001) Year of the Fires – the Story of the Great Fires of 1910. Viking Press, New York.
Ramdass, I. and Haniff, M. (1990) A Definition of Priority Conservation Areas in Amazonia. Guyana country
paper. Biological priorities for conservation in Amazonia. Conservation International, Washington,
DC.
Räsänen, M.E., Salo, J.S. and Kalliola, R.J. (1987) Fluvial perturbance in the western Amazon basin: regula-
tion by long-term sub-Andean tectonics. Science 238, 1398–1401.
Redford, K.H. (1993) Hunting in neotropical forests: a subsidy from nature. In: Hladik, C.M., Hladik, A.,
Linares, O.F., Pagezy, H., Sample, A. and Hadley, M. (eds) Tropical Forests, People and Food.
Parthenon Publishing, Paris, pp. 227–246.
Repetto, R. and Gillis, M. (eds) (1988) Public Policies and the Misuse of Forest Resources. Cambridge
University Press, Cambridge, UK.
Richards, P.W. (1952) The Tropical Rain Forest. Cambridge University Press, Cambridge, UK.
Richards, P.W. (1996) The Tropical Rain Forest: An Ecological Study. Cambridge University Press,
Cambridge, UK.
Rodríquez, J.P. and Rojas-Suárez, F. (1996) Guidelines for the design of conservation strategies for the ani-
mals of Venezuela. Conservation Biology 10, 1245–1252.
Rollet, B. (1980) Jari: succès ou échec? Un exemple de développement agro-sylvo-pastoral et industriel en
Amazonie brésilienne. Bois et Forêts des Tropiques 192, 3–34.
Roulet, M., Lucotte, M., Farella, N., Serique, G., Coelho, H., Sousa Passos, C.J., Jesus da Silva, E., Scavone
de Andrade, P., Mergler, D., Guimaraes, J.R.D. and Amorim, M. (1999) Effect of recent human colo-
Forest Conservation and Management 519

nization on the presence of mercury in Amazonian ecosystems. Water, Air and Soil Pollution 112,
297–313.
Russell, C.E. (1987) Plantation forestry. In: Jordan, C.F. (ed.) Amazonian Rain Forests – Ecosystem
Disturbance and Recovery. Springer, Berlin, pp. 76–89.
Santos, G.M., Cordeiro, R.C., Silva Filho, E.V., Turcq, B., Lacerda, L.D., Fifield, L.K., Gomes, P.R.S.,
Hausladen, P.A., Sifeddine, A. and Albuquerque, A.L.S. (2001) Chronology of the atmospheric mercury
in Lagoa da Pata basin, upper Rio Negro region of Brazilian Amazon. Radiocarbon 43, 801–808.
Schulz, J.P. (1960) Ecological studies on rain forest in Northern Suriname. Verhandelingen der Koninklijke
Nederlandse Akademie van Wetenschappen, Afd. Natuurkunde 163, 1–250.
Silva, J.M.N., de Carvalho, J.O.P., Lopes, J.C.A., de Almeida, B.F., Costa, D.H.M., de Oliviera, L.C., Vanclay,
J.K. and Skovs.gaard, J.P. (1995) Growth and yield of a tropical rain forest in the Brazilian Amazon 13
years after logging. Forest Ecology and Management 71, 267–274.
Silva-Forsberg, M.C., Forsberg, B.R. and Zeidemann, V.K. (1999) Mercury contamination in humans linked
to river chemistry in the Amazon Basin. Ambio 28, 519–521.
Sing, K.A., Hryhorczuk, D.O., Saffirio, G., Sinks, T., Paschal, D.C. and Chen, E.H. (1996) Environmental
exposure to organic mercury among the Makuxi in the Amazon basin. International Journal of
Occupational and Environmental Health 2, 165–171.
Sist, P., Nolan, T., Bertault, J.-G. and Dykstra, D. (1998) Logging intensity versus sustainability in Indonesia.
Forest Ecology and Management 108, 251–260.
Spangenberg, A., Grimm, U., Sepeda da Silva, J.R. and Fölster, H. (1996) Nutrient store and export rates of
Eucalyptus urograndis plantations in eastern Amazonia (Jari). Forest Ecology and Management 80,
225–234.
Spencer, C.S., Redmond, E.M. and Rinaldi, M. (1998) Prehispanic causeways and regional politics in the
llanos of Barinas, Venezuela. Latin American Antiquity 9, 95–110.
Stattersfield, A.J., Crosby, M.J., Long, A.J. and Wege, D.C. (1998) Endemic Bird Areas of the World – Priorities
for Biodiversity Conservation. Birdlife Conservation Series No. 7. Birdlife International, London.
ter Steege, H. (1998) The use of forest inventory data for a National Protected Area Strategy in Guyana.
Biodiversity Conservation 7, 1457–1483.
ter Steege, H. (2000) Plant Diversity in Guyana. Tropenbos Series 18. Tropenbos Foundation, Wageningen,
The Netherlands.
ter Steege, H. and Hammond, D.S. (1996) Forest management in the Guianas: ecological and evolutionary
constraints on timber production. BOS Nieuwsletter 15, 62–69.
ter Steege, H. and Hammond, D.S. (2001) Character convergence, diversity, and disturbances in tropical rain
forest in Guyana. Ecology 82, 3197–3212.
ter Steege, H. and Zondervan, G. (2000) A preliminary analysis of large-scale forest inventory data of the
Guiana Shield. In: ter Steege, H. (ed.) Plant Diversity in Guyana. Tropenbos Foundation, Wageningen,
The Netherlands, pp. 35–54.
ter Steege, H., Jetten, V.G., Polak, A.M. and Werger, M.J.A. (1993) Tropical rain forest types and soil factors
in a watershed area in Guyana. Journal of Vegetation Science 4, 705–716.
ter Steege, H., Boot, R.G.A., Brouwer, L.C., Hammond, D.S., van der Hout, P., Jetten, V.G., Khan, Z., Polak,
A.M., Raaimakers, D. and Zagt, R. (1995) Basic and applied research for sound rain forest management
in Guyana. Ecological Applications 5, 904–910.
ter Steege, H., Sabatier, D., Castellanos, H., van Andel, T., Duivenvoorden, J.F., de Oliveira, A.A., Ek, R.C.,
Lilwah, R., Maas, P.J.M. and Mori, S.A. (2000) An analysis of Amazonian floristic composition, includ-
ing those of the Guiana Shield. Journal of Tropical Ecology 16, 801–828.
ter Steege, H., Welch, I.A. and Zagt, R. (2002) Long-term effect of timber harvesting in the Bartica Triangle,
Central Guyana. Forest Ecology and Management 170, 127–144.
Thompson, J., Proctor, J., Viana, V., Milliken, W., Ratter, J.A. and Scott, D.A. (1992) Ecological studies on a
lowland evergreen rain forest on Maracá Island, Roraima, Brazil. I. Physical environment, forest struc-
ture and leaf chemistry. Journal of Ecology 80, 689–703.
USGS and CVGTM (1993) Geology and Mineral Resource Assessment of the Venezuelan Guayana Shield.
USGS Bulletin 2062. US GPO, Washington, DC.
van Andel, T. (1998) Commercial exploitation of non-timber forest products in the north-west district of
Guyana. Caribbean Journal of Agriculture and Natural Resources 2, 15–28.
van Andel, T. (2000) Non-timber Forest Products of the North West District of Guyana. Part I. Tropenbos-
Guyana Series 8A. Tropenbos-Guyana Programme, Georgetown.
van der Hout, P. (1999) Reduced Impact Logging in the Tropical Rain Forest of Guyana. Ecological,
520 D.S. Hammond

Economic and Silvicultural Consequences. Tropenbos-Guyana Series 6. Tropenbos-Guyana


Programme, Georgetown.
van der Hout, P. (2000) Testing the applicability of reduced impact logging in greenheart forest in Guyana.
International Forestry Review 2, 24–32.
Vanclay, J. (2001) The effectiveness of parks. Science 293, 1007.
Veiga, M.M., Meech, J.A. and Onate, N. (1994) Mercury inputs from forest fire in the Amazon. Nature 368,
816–817.
Veillon, J.P. (1985) El crecimiento de algunas bosques naturales de Venezuela en relación con los parámet-
ros del medio ambiente. Revista Forestal Venezolana 29, 5–121.
Vink, A.T. (1970) Forestry in Surinam – a Review of Policy, Planning, Projects, Progress. Suriname Forest
Service (LBB), Paramaribo.
Welch, I.A. (1975a) A Short History of the Guyana Forest Department 1925–1975. Guyana Forestry
Commission, Georgetown.
Welch, I.A. (1975b) The Timber Resources of Guyana. Technical report. Guyana Forest Department,
Georgetown.
Whiteman, A. (1999) Economic Rent from Forest Operations in Suriname and a Proposal for Revising its
Forest Revenue System. FAO, Rome.
Williams, D. (1994) Annex C. Archaeology and anthropology. In: Environmental Impact Assessment – the
Linden Lethem Road. ERM, Georgetown, Guyana
Winkler, N. (1997) Environmentally Sound Forest Harvesting – Testing the Applicability of the FAO Model
Code in the Amazon in Brazil. Forest Harvesting Case Study No. 8. FAO, Rome.
Zagt, R.J. (1997) Tree Demography in the Tropical Rain Forest of Guyana. Tropenbos-Guyana Series 3.
Tropenbos-Guyana Programme, Georgetown.

You might also like