01.06.2013 Views

Untitled - University of Oxford

Untitled - University of Oxford

Untitled - University of Oxford

SHOW MORE
SHOW LESS

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

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

OFI OCCASIONAL PAPERS<br />

No 49<br />

The Natural Management <strong>of</strong> Tropical Forests<br />

for Timber and Non-Timber Products<br />

Sarah Laird<br />

1995<br />

ISBN 0 85074 136 X<br />

ISSN 0269 - 5790<br />

<strong>Oxford</strong> Forestry Institute<br />

Department <strong>of</strong> Plant Sciences<br />

<strong>University</strong> <strong>of</strong> <strong>Oxford</strong>


Table <strong>of</strong> Contents<br />

Acknowledgements ii<br />

I. Introduction 1<br />

11. Sustainable Forestry 2<br />

Ill. The Economic Underpinnings <strong>of</strong> Natural Forest Management 4<br />

- Box 1: Lesser-Known Timber Species 9<br />

- Box 2: The Economic Valuation <strong>of</strong> Non-Timber Forest Products 10<br />

IV. The Ecologically Sustainable Harvest <strong>of</strong> Non-Timber Products<br />

from Tropical Forests 11<br />

- Plant Parts Used in Non-Timber Forest Products 14<br />

- Wildlife as a Non-Timber Forest Product 17<br />

V. The Ecological Underpinnings <strong>of</strong> Natural Forest Management 18<br />

VI. Harvesting Operations in Managed Natural Forest 20<br />

- Box 3: Inventories <strong>of</strong> Non-Timber Forest Products for<br />

Natural Forest Management 25<br />

VII. Silvicultural Systems for Natural Forest Management 26<br />

- Box 4: Examples <strong>of</strong> MUlti-purpose Amazonian Species with<br />

Present and Future Potential for Natural Forest Management 33<br />

VIII. Traditional Management <strong>of</strong> Neotropical Forests for Timber and<br />

Non-Timber Products 35<br />

IX. Natural Forest Management and the Conservation <strong>of</strong> Biodiversity 38<br />

- The Impact <strong>of</strong> Natural Forest Management on Wildlife 42<br />

x. Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 45<br />

Bibliography 49<br />

Page


ii<br />

Acknowledgements<br />

I would like to thank the following for their valuable comments, advice and thoughts on this<br />

document: Trish Shanley, Mike Arnold, Nick Brown, and Peter Kanowski.


I. Introduction<br />

The tenn "non-timber forest products" (NTFPs) describes a wide range <strong>of</strong> products, including<br />

medicinal plants, fibres, resins, latexes, fruits, foods, construction materials, and bushmeat. The plant<br />

products may be taken from a variety <strong>of</strong> life forms and plant parts, including reproductive propagules,<br />

plant exudates, vegetative structures, roots and bark. NTFPs may be harvested for subsistence<br />

consumption, for small local markets, or for regional or even international trade. They may have long<br />

histories in traditional resource use, or they may be recently developed, and have applications <strong>of</strong><br />

primary importance to industrial cultures. NTFPs may be sourced from high-diversity primary forest,<br />

low diversity "oligarchic" forests, or secondary forests. These forests may be in various states <strong>of</strong><br />

repair, and may exist in extractive reserves, indigenous reserves, communally-held and managed<br />

forests, privately-owned lands, reserves demarcated for conservation purposes, timber concessions,<br />

or government lands. NTFPs may be harvested from areas with high population densities with welldeveloped<br />

labour and trade patterns, or in areas that are sparsely populated. Harvesters may be rural<br />

peoples (in which case they might be indigenous, long-term settlers <strong>of</strong> mixed ethnicity, or recent<br />

immigrants from other rural areas or urban centres), or teams <strong>of</strong> individuals from urban areas. These<br />

harvesters mayor may not operate under traditional systems <strong>of</strong> resource management and harvesting<br />

controls. Products might be consumed in rural areas or urban centres. Trade patterns may be complex,<br />

with a variety <strong>of</strong> intermediaries, or fairly simple. This is all to say that when we speak <strong>of</strong> "non-timber<br />

forest products" (a term which in itself has been questioned, due to the implied primacy <strong>of</strong> timber<br />

forest products), we are speaking <strong>of</strong> many things (Padoch et al., 1992; Pendleton, 1992; Vasquez and<br />

Gentry, 1989).<br />

Timber production can have a number <strong>of</strong> effects on the production <strong>of</strong> non-timber forest products.<br />

Destructive harvesting operations can cause direct damage to species in residual stands and those that<br />

make up the understorey and ground cover <strong>of</strong> forests, many <strong>of</strong> which are important NTFPs.<br />

Subsequent silvicultural treatments can reduce species diversity by promoting an increased proportion<br />

<strong>of</strong> commercial species, and removing the competing "undesirables", many <strong>of</strong> which are important<br />

NTFPs. In some cases, timber species will have important non-timber uses locally, and the overharvesting<br />

<strong>of</strong> these species for timber will reduce collection <strong>of</strong> these species for their non-timber uses.<br />

The harvest <strong>of</strong> NTFPs could, at times, reduce the number and regeneration success <strong>of</strong> timber species,<br />

as well. For example, if bark is stripped, reproductive propagules over-harvested, or exudates tapped<br />

heavily, growth rates and reproduction <strong>of</strong> timber species can be impaired.<br />

Timber harvesting and silvicultural treatments, by reducing the structural and species diversity <strong>of</strong> a<br />

forest, can also produce a number <strong>of</strong> still largely unknown ecological repercussions. These result in<br />

reductions in numbers <strong>of</strong> pollinators, seed dispersers, and alterations in plant-herbivore relationships,<br />

and may ultimately produce conditions in which it is difficult for many forest species to regenerate.<br />

A wide range <strong>of</strong> NTFPs are usually harvested from a given forest, so reductions in species diversity<br />

can directly impact the collection <strong>of</strong> these products, and the consumption and trade patterns <strong>of</strong> local<br />

people dependent upon them for their livelihoods and survival.<br />

Timber and non-timber products can be incorporated into multi-purpose systems <strong>of</strong> natural forest<br />

management that both minimize the negative impacts <strong>of</strong> timber extraction and capitalize on the many<br />

benefits provided by a range <strong>of</strong> forest products. Following is a discussion <strong>of</strong> some <strong>of</strong> the economic,<br />

cultural and ecological issues involved in this type <strong>of</strong> system, with particular emphasis on the<br />

Neotropics, and general suggestions for guidelines that might assist with the integration <strong>of</strong> non-timber<br />

and timber products in forest management.<br />

1


2<br />

11. Sustainable Forestry<br />

Duncan Poore (1989), in his popularly quoted definition <strong>of</strong> sustainable timber production, describes<br />

"the single most important condition to be met is that nothing should be done that will irreversibly<br />

reduce the potential <strong>of</strong> the forest to produce marketable timber - that is, there should be no<br />

irreversible loss <strong>of</strong> soil, soil fertility or genetic potential <strong>of</strong> marketable species". Sustained yield<br />

timber production implies the management <strong>of</strong> timber crops so as to produce a continuous flow <strong>of</strong><br />

wood output, while safeguarding the productive potential <strong>of</strong> the forest (Gane, 1992). Sustainable<br />

forestry implies a much broader, multi-sectorial agenda that attempts to "balance use", includes the<br />

production <strong>of</strong> non-timber as well as timber products, conserves forest ecosystems and the genetic and<br />

species diversity contained within them, and generates income, employment, consumption, and<br />

investment benefits for local forest dwellers, as well as nations (Gane, 1992; Poore, 1989; Barbier,<br />

1987).<br />

While there is a general agreement that ecological and technical abilities are sufficient to achieve<br />

sustainable timber production, it could be questioned whether they are up to the job <strong>of</strong> sustainable<br />

forestry, as described above. The complexities <strong>of</strong> natural forest management appear to only multiply<br />

as more knowledge is acquired, and the shifting <strong>of</strong> priorities from timber production to multiple use<br />

natural forest management carries with it an enormous burden <strong>of</strong> ignorance. Not only is little known<br />

about the ecological requirements <strong>of</strong> individual commercial timber species, less still is understood<br />

about non-timber commercial species.<br />

The Supply <strong>of</strong> Tropical Timber<br />

Global demand for industrial wood has been projected to grow by margins <strong>of</strong> 15% to 40% over 15<br />

years, and from 35% to 75% over 50 years (Arnold, 1991). Tropical hardwoods from Southeast Asia<br />

currently fonn the bulk <strong>of</strong> the export market from developing countries, accounting for 84% <strong>of</strong> the<br />

market in 1981; 14% <strong>of</strong> these exports come from Africa, and 1% from Latin America (although<br />

forests are logged heavily in Africa and Latin America for regional consumption (Prescott-Allen,<br />

1982; Schmidt, 1991; Repetto and Gillis, 1988)1. Of this, the FAO found in 1985 that over 90%<br />

came from unmanaged natural forest, roughly divided between primary and once-logged forest'. Of<br />

the 800,000,000 ha <strong>of</strong> forests in IITO producer member countries, only 1,000,000 ha are managed<br />

under sustained yield (Poore, 1989). In the Neotropics, for every 1 ha <strong>of</strong> tropical moist forest<br />

managed under sustained yield management, 35,000 ha are not (Wadsworth, 1987).<br />

Although plantations will play an increasing role in timber production in tropical countries, alone they<br />

will not be able to satisfy the increasing demand for wood. Plantations must also compete with<br />

agriculture for valuable land and must be near processing industries; they tie up capital - <strong>of</strong> which<br />

there is <strong>of</strong>ten little in these regions; they are vulnerable to attacks by pests and diseases, and are<br />

generally perceived as a risk to investors (Wadsworth, 1987; Schmidt, 1991). Natural forest<br />

1 Regional markets <strong>of</strong>ten consume the bulk. <strong>of</strong> a country's timber. For example, in Brazil, 80% <strong>of</strong> the wood produced<br />

in Amazonia is consumed by regional markets, with national and international markets making up the remaining 20% (UbI<br />

et ai., 1989).<br />

2 Rotations, regeneration periods, felling cycles, harvestable girth limits, etc. were not based on growth rates and<br />

regeneration requirements <strong>of</strong> the species, but on the demand for wood, so forests are usually harvested in excess <strong>of</strong> the<br />

allowable cut, and logging damage tends to be severe (Nair, 1991).


management and plantations <strong>of</strong> tropical timber, therefore, must play complementary roles in the<br />

production <strong>of</strong> tropical timber (Schmidt, 1991; Anderson, 1990).<br />

Natural Forest Management<br />

Many <strong>of</strong> the problems associated with natural forest managemenf result from the limited objectives<br />

set in the past by forest managers 4 • The best way to ensure success in natural tropical forests is to<br />

use as much <strong>of</strong> the great species and genetic diversity as possible (Gomez-Pompa and Burley, 1991;<br />

Hart, 1980; Clay and Clement, 1993; Gentry, 1992; Hidalgo, 1992). The diversity <strong>of</strong> tropical forests<br />

has been seen in the past as a major obstacle to its commercial exploitation, and numerous forest<br />

management efforts have failed because they fought against this diversity, rather than capitalized on<br />

it. But natural forests managed for multiple purposes have a number <strong>of</strong> significant advantages: they<br />

produce a variety <strong>of</strong> products, and so can flexibly adjust to changing markets for forest products; the<br />

costs <strong>of</strong> management are significantly lower than more intensive land-use systems; there is less<br />

likelihood <strong>of</strong> pests, disease, or natural disasters destroying their productive capacity; they are more<br />

easily adapted to local social and cultural conditions; and they maintain the ecological functions <strong>of</strong><br />

the forest (Poore and Sayer, 1991; Anderson, 1990). Ultimately, natural forest management provides<br />

the best compromise between the need to use valuable forest resources and the need to conserve<br />

forests and the diversity <strong>of</strong> species they contain (ITIO, 1992; Salick, 1992). As Schmidt (1991) said,<br />

"is natural forest management uneconomic or is it the only economic alternative?".<br />

Standing tropical forests provide numerous non-timber benefits, including climatic and environmental<br />

services, wild relatives <strong>of</strong> crops, potential phannaceutical compounds 5 , and numerous non-timber<br />

forest products. Non-timber products like medicines, resins, essential oils, fibres) and foods, generate<br />

large amounts <strong>of</strong> income in the local and regional markets <strong>of</strong> tropical countries, and some species,<br />

like rattan, rosewood oil, chicle, and Brazil nuts, are traded widely on the international market 6 • The<br />

non-timber forest products <strong>of</strong> three areas in Latin America, for example, yielded greater potential<br />

returns!hectare than timber extraction, cattle ranching, and plantation forestry (Peters et al., 1989;<br />

Balick and Mendelssohn, 1992; Grimes et al., 1993). More than 1.5,000,000 people in the Brazilian<br />

Amazon earn the bulk <strong>of</strong> their living from the extraction <strong>of</strong> forest products (Richards, 1993). The<br />

3 Natural forest management can be defined as the extraction <strong>of</strong> some <strong>of</strong> the forest products without significantly<br />

disturbing the ecosystem as a whole. Forest manipulation and conservation occur through an efficient natural or induced<br />

regeneration <strong>of</strong> the forest ecosystem, which is only lightly disturbed (Gomez-Pompa and Burley, 1991; Schmidt, 1991).<br />

The sustainable implementation <strong>of</strong> this system results in the maintenance <strong>of</strong> ecological processes and significant levels<br />

<strong>of</strong> biodiversity, and requires the sustainable harvest <strong>of</strong> non-timber and timber products alike (poore and Sayer, 1991).<br />

4 The failure <strong>of</strong> sustainable natural forest management has been attributed to a number <strong>of</strong> factors, including: high<br />

extraction costs (in relation to prices) associated with selective logging; low volume commercial woods available/unit area,<br />

and so low economic returns; lack <strong>of</strong> understanding <strong>of</strong> the silvics <strong>of</strong> tropical tree species, and resulting difficulties in<br />

management; vulnerability to disruptive land uses, such as shifting cultivation; government policies that make sustainedyield<br />

forestry economically unattractive; and forestry agencies that don't effectively regulate forestry practices (UbI et<br />

al., 1989; Anderson, 1990).<br />

s Principe (1992) estimates that the value for prescription and over-the-counter medicines in OEeD countries in 1985<br />

was $43,000,000,000. Additionally, the World Health Organization estimates that the vast majority (80%) <strong>of</strong> the world's<br />

population depends upon traditional, natural product-based medicines for their primary healthcare.<br />

6 Exports <strong>of</strong> non-timber forest products from Indonesia reached $125,OOO,OOO/year in the early 1980s. In 1973, nonwood<br />

products accounted for 2.9% <strong>of</strong> the total forest product export value, but in 1982 had increased to 13.3%. The rattan<br />

industry alone employs 70,000 - 100,000 people, not including collectors (De Jong and Mendelssohn, 1992).<br />

3


4<br />

yields on alternative land-uses for the tropics have been greatly overestimated. Poor soils and<br />

invasions by pests and weeds have led to the failure <strong>of</strong> most intensive land-use systems (Peters et<br />

al., 1989; Balick and Mendelssohn, 1992). Natural forests can be managed to efficiently utilize the<br />

wide variety <strong>of</strong> possible forest resources, providing the maximum total benefit over the long run, and<br />

avoiding the environmental degradation associated with many other land-use options in the tropics<br />

(Repetto and Gillis, 1988).<br />

Ill. The Economic Underpinnings <strong>of</strong> Natural Forest Management<br />

Economics is a way <strong>of</strong> thinking about problems, not a recipe for solving them - Keynes, 1936<br />

Economic theory suffers from a tendency to assume that the world is a sum <strong>of</strong> its parts - that if we<br />

objectively understand the many components <strong>of</strong> a system, we will understand the whole. But in<br />

economics, as in ecology, the reality is far more complex and there exist no "universal truths"; the<br />

"parts" represent constantly co-evolving, dynamic and interdependent relationships, difficult to<br />

understand and impossible to separate from each other in any applicable analysis (Norgaard).<br />

Neoclassical economic theory has for decades determined the type <strong>of</strong> forestry and "development"<br />

practised in the tropics, promoting national capital accumulation as a measure <strong>of</strong> a linear type <strong>of</strong><br />

"progress".<br />

Microeconomic models utilizing comparative advantage, specialization, and gains from exchange<br />

were employed to increase GNP and per capita incomes, but little "real" development resulted.<br />

Although overall growth was achieved in many countries, it was <strong>of</strong>ten at the expense <strong>of</strong> natural<br />

resource wealth and resulted in only wider gaps between rich and poor (Schwartzman, 1992; Burns,<br />

1986; Barbier, 1987). The concept <strong>of</strong> development, much as with "sustainability", has been revised<br />

to emphasize meeting the basic needs <strong>of</strong> the poor, to advocate cultural sensitivity, and to encourage<br />

"grassroots" participation in the development process. In theory, this means that political, social and<br />

cultural values are added to measurements <strong>of</strong> economic activity when assessing development, and<br />

smaller, adaptive projects are encouraged (Barbier, 1987).<br />

Neoclassical economic theory holds that resources will be efficiently allocated when prices reflect<br />

true resource scarcity, there exists a right <strong>of</strong> ownership and therefore freedom to trade, and when<br />

consumers have access to information about the total amount <strong>of</strong> a resource available (National<br />

Research Council, 1992). But tropical forest resources have consistently been undervalued, and their<br />

over-exploitation has been considered net income, rather than a loss <strong>of</strong> wealth (McNeely, 1988). As<br />

a result, under this system the only biodiversity remaining will be that which is too expensive to<br />

exploit, due to difficulties in access, and too immediately valuable to destroy, such as ground water.<br />

The failure in this system to value biodiversity results from a combination <strong>of</strong> market failures,<br />

intervention failures, and global appropriations failures (Pearce, 1993), including: biodiversity<br />

provides many non-marketed environmental services (like erosion prevention and flood control), and<br />

supplies "invisible" products (like NTFPs traded only on local markets), and many potential products<br />

(like lesser-known-timber species); government interventions, such as (perverse) incentives, tax<br />

provisions, and tax credits, distort commodity prices and promote short-tenn pr<strong>of</strong>its through resource


depletion 7 ; biodiversity and many forests are a "public good" and therefore are not "owned", so can<br />

result in infinite discounting <strong>of</strong> future costs on the part <strong>of</strong> users; consumers do not have adequate<br />

access to infonnation about the total value <strong>of</strong> natural resources and so prices do not reflect scarcity;<br />

timber and some other forest products have long gestation periods, and so the discount rates applied<br />

by current economic planners make its long-term sustainable use appear uneconomic; knowledge <strong>of</strong><br />

the multiplicity <strong>of</strong> potential useful forest species is limited, and so not factored into assessments <strong>of</strong><br />

its value; and forest products are most widely used by the poor and politically and economically<br />

powerless (McNeely, 1988; National Research Council, 1992; Pearce, 1992; Barbier, 1987;<br />

Schwartzman, 1992; Repetto and Gillis, 1988; Swanson, 1992; Repetto and Gillis, 1988; Uhl et al.,<br />

1989).<br />

Causal mechanisms also influence the use <strong>of</strong> forest resources. These might include economic<br />

instruments used to stimulate the economy, property rights 8 , external debt, international trade<br />

policies, the over-valuation <strong>of</strong> currency, inflation (which encourages cattle ranching, for example,<br />

because cattle are - in effect - inflation-pro<strong>of</strong>), price instability <strong>of</strong> major export crops, the structure<br />

(time and cost) <strong>of</strong> concessions, etc. (National Research Council, 1992; Repetto and Gillis, 1988)9.<br />

Direct economic incentives and disincentives can also influence the nature <strong>of</strong> resource use.<br />

Disincentives for destruction include the posting <strong>of</strong> environmental bonds by logging companies;<br />

incentives might include the employment <strong>of</strong> local people in reserved areas (McNeely, 1988).<br />

Proper valuation techniques are necessary for the many timber and non-timber values generated by<br />

tropical forests, and for the appraisal <strong>of</strong> forestry projects and alternative forest uses. Assessments <strong>of</strong><br />

the financial value <strong>of</strong> various forest management systems does not even roughly approximate the<br />

wider economic value <strong>of</strong> all benefits (irrespective <strong>of</strong> who receives them), against all the costs<br />

(irrespective <strong>of</strong> who bears them) (Leslie, 1987). Foresters have been forced to increase the fmancial<br />

perfonnance <strong>of</strong> natural forest management by increasing revenues and decreasing costs; silvicultural<br />

systems strive to increase growth rates and yield, shorten rotations and lower treatment and protection<br />

costs; lesser-known species are sought to increase yield per unit area (Leslie, 1987). But forests<br />

provide numerous benefits and excessive exploitation incurs many costs not measured by financial<br />

analysis.<br />

A number <strong>of</strong> economic valuation techniques have been developed to incorporate non-marketed goods<br />

and services provided by forests, so as to be comparable with values placed on marketable goods.<br />

These include costs <strong>of</strong> replacement, hedonic price analysis, travel cost methods, and contingent<br />

valuation. These are used to measure both direct and indirect use <strong>of</strong> a resource (timber, plant<br />

7 For example, in the Brazilian Amazon subsidies and other policy distortions are estimated to have accounted for<br />

at least 30% <strong>of</strong> all forest altered by 1980. Over the past three decades> 10,000,000 ha <strong>of</strong> Amazonian rainforest have been<br />

converted to cattle pastures, largely through government policies. For every 1/4 lb. hamburger that cost US$ 0.26 to<br />

produce, the government <strong>of</strong> Brazil expended US$ 0.22 (Anderson, 1990). From 1970-1988, US$23,OOO,OOO,OOO was spent<br />

subsidizing cattle ranching, agriculture and colonization projects, and the building <strong>of</strong> roads in the Amazon (Schwartzman,<br />

1992).<br />

8 Private property and communal property rights best protect forest resources in the long-term. Open-access and<br />

insecure land tenure, because they do not guarantee the user any future access, usually results in a "tragedy <strong>of</strong> the<br />

commons" in which the future is discounted infmitely and user costs are ignored.<br />

9 Mather (1990) suggest that policies for land use systems other than forestry have had a larger impact on forests than<br />

any forestry policy.<br />

5


6<br />

breeding, landscape, recreation, biodiversity, watershed, recreation, etc.), option values (biodiversity,<br />

recreation, potential phannaceuticals, landscape, etc.) and existence values (biodiversity, landscape,<br />

etc.) (Pearce, 1991; 1992).<br />

Standard cost-benefit analysis <strong>of</strong> the variety <strong>of</strong> products produced by a forest can also be used in<br />

some cases to provide strong arguments for sustainable use, although Leslie (1987) and others find<br />

it "tactically" naive to fight the battle for natural forest management on financial grounds (National<br />

Research Council, 1992; Pearce, 1991, 1992; Leslie, 1987; ITTO, 1990) But because most policymakers<br />

use cost-benefit analysis when making decisions about land-use options, it is important that<br />

these be expanded to include the range <strong>of</strong> marketable products, including NTFPs. However, because<br />

the trade in NTFPs is largely local or regional and takes place in the "infonnal sector", few records<br />

exist on their economic value; as a result, although NTFPs make up a significant percentage <strong>of</strong> the<br />

actual value <strong>of</strong> the forests, they are not figured in to policy and management decisions (Padoch et<br />

al., 1992; Cleary, 1992).<br />

Comparative economics has been used to successfully demonstrate the economic importance <strong>of</strong><br />

conservation and multi-purpose use <strong>of</strong> natural forests, by demonstrating the superior economic<br />

perfonnance <strong>of</strong> NTFPs when compared with other land-use options in tropical forests (Peters et al.,<br />

1989; Balick and Mendelssohn, 1992; Godoy and Lubowski, 1992; Falconer, 1990; De Beer and<br />

McDermott, 1989; Phillips, 1992; Grimes et al., 1993). Additionally, the harvest <strong>of</strong> NTFPs throughout<br />

the economically "unproductive" building phase <strong>of</strong> timber trees can subsidize silvicultural and<br />

management expenses, and can lessen the blow struck to financial analyses <strong>of</strong> natural forest<br />

management by compounded interest 10 • For local peoples, non-timber products can provide<br />

consistent, if relatively small, sources <strong>of</strong> income, complemented by the periodic concentration <strong>of</strong><br />

capital produced by the felling <strong>of</strong> timber (Leslie, 1987; Salick, 1992; Reining and Heinzman, 1992;<br />

Malhotra et al., 1991).<br />

But valuation is clearly not enough. Peters et al. (1989) note that, although NTFPs have a superior<br />

economic value in the area <strong>of</strong> Peru they studied, timber continues to be considered <strong>of</strong> primary<br />

importance, largely due to its sale on international markets and generation <strong>of</strong> foreign exchange.<br />

NTFPs, on the other hand, are collected and sold in local markets by an incalculable number <strong>of</strong><br />

subsistence farmers, forest collectors, middlemen and shop owners. These decentralized trade<br />

networks are extremely hard to monitor and easy to ignore (Peters et al., 1989; Padoch, 1992).<br />

Some Socioeconomic Considerations for the Sustainable Use <strong>of</strong> NTFPs<br />

In order for non-timber forest products to be incorporated into natural forest management and<br />

generate sustainable income for local peoples, a number <strong>of</strong> basic issues must be addressed ll . These<br />

include: land tenure, fluctuations in markets, local value-added processing, over-harvesting <strong>of</strong> species,<br />

transportation, and the expansion <strong>of</strong> markets overseas.<br />

10 For example, rattan species have been planted in logged-over forest in Southeast Asia to both "rehabilitate"<br />

degraded forests and provide a significant interim income while the forest recovers from logging pressure (Sayer, 1991).<br />

11 Ifnatural forest management is to succeed, forest managers must better understand the ideological assumptions from<br />

which their decisions are made, and correct deficiencies in their understanding <strong>of</strong> the socioeconomic context in which<br />

forestry projects operate. This is particularly important when incorporating NTFPs into forest management, because<br />

ignorance <strong>of</strong> the socioeconomic dynamics <strong>of</strong> these complex collection and trading systems could lead to projects that<br />

would restrict and potentially damage local livelihoods, rather than enhance the productivity <strong>of</strong> a forest (Cleary, 1992).


Land Tenure<br />

Land tenure and resource rights for forest residents are essential prerequisites to sustainability in an<br />

economy based on natural forest management that utilizes a wide range <strong>of</strong> products over long periods<br />

<strong>of</strong> time (Clay and Clement, 1993; Schwartzman, 1992; Allegreti, 1990; Cleary, 1992; De long and<br />

Mendelssohn, 1992). Because <strong>of</strong> the scattered nature <strong>of</strong> most species in the tropical forest, and the<br />

low productivity per unit area, sustainable, lower-yielding systems are only attractive if collectors<br />

retain long-tenn security <strong>of</strong> access to the forest and its resources (Phillips, 1993).<br />

Unpredictable Markets<br />

External markets for NTFPs tend to be unpredictable, operating on boom-bust cycles, with synthetics<br />

substituting for natural products once an economy <strong>of</strong> scale is achieved (Padoch, 1992; Richards,<br />

1993; Sizer, 1992). For example, during the early part <strong>of</strong> this century, Brazil exported 40 different<br />

vegetable oils from the Amazon. The spread <strong>of</strong> electricity (and so decreased reliance on candles) and<br />

extensive cultivation <strong>of</strong> corn and soybeans (which became the most popular vegetable oils<br />

internationally) resulted in the decline <strong>of</strong> markets for these oils (Clay and Clement, 1993). But today,<br />

there exists an increasing global market for vegetable oils, especially exotic, on the part <strong>of</strong>companies<br />

looking for environmentally-sound substitutes for plantation-grown palm oil. Chicle (used in chewing<br />

gum) and tagua (used to make buttons) were both displaced by synthetic substitutes, but have recently<br />

regained popularity through expanded marketing ventures in the United States (Reining and<br />

Heinzman, 1992; Ziffer, 1992).<br />

By maintaining a diversity <strong>of</strong> products in the extractive economy, rather than depending upon a few,<br />

the potential impacts <strong>of</strong> individual product price fluctuations on the value <strong>of</strong> the forests for all NTFPs<br />

can be minimized (Grimes et al., 1993; Clay, 1992; Clay and Clement, 1993). Markets for a single<br />

NTFP species can also be diversified. For example, Brazil nut (Bertholletia excelsa) can be used as<br />

nuts, in ice cream, baked goods, candy, oil, and as a flour (Clay and Clement, 1993).<br />

It is important to note, however, that the bulk <strong>of</strong> NTFPs are traded on local or domestic markets.<br />

While some <strong>of</strong> these markets might decline or disappear over time, either because the product is an<br />

"inferior good" that drops out <strong>of</strong> consumption patterns as incomes rise, or because <strong>of</strong> competition<br />

from factory-made alternatives, much <strong>of</strong> this trade is in products which enjoy stable or growing<br />

markets (Arnold, pers. comm., 1994).<br />

Adding Value Locally to NTFPs<br />

Adding value to products destined for markets can be an important way to increase benefits for local<br />

forest residents. Because adding value to forest products involves the combined efforts <strong>of</strong> collectors,<br />

producers and marketers that do not usually work together, these efforts can be plagued by the<br />

politics <strong>of</strong> ownership and control over infrastructure and processing facilities (Clay and Clement,<br />

1993; Clay, 1992). The quality <strong>of</strong> products produced through local processing must be comparable<br />

with that produced elsewhere, or these products will not sell on the open market.<br />

Expanding the Markets Overseas<br />

Wild plants and animals are subject to a management "Catch 22": if their economic utility is<br />

overlooked or ignored, or if their use is in competition with another form <strong>of</strong> land use, they may<br />

suffer from habitat destruction; if, however, their economic utility is evident, they are likely to be<br />

over-exploited (Prescott-Allen, 1982). In building up markets for NTFPs located in a forest, a number<br />

<strong>of</strong> international marketers <strong>of</strong> NTFPs and researchers suggest that one should begin with existing<br />

products, rather than attempt to develop new ones, ideally replacing an existing commodity with<br />

something superior or cheaper, or, in some cases, supplying unfilled demand (FAO, 1991; Gentry,<br />

7


8<br />

1992; Ziffer, 1992; Clay, 1992). These should be those products with the highest market price, the<br />

most reliable supply, and the greatest potential for future market expansion (Peters, 1993).<br />

"Green" markets in developed countries for sustainably-produced NTFPs appear to be significant and<br />

growing. For example, in 1991 39% <strong>of</strong> the American public considered themselves environmentalists<br />

and 50% said that they would select a product for purchase based on environmental concerns (Dixon<br />

et al., 1991; Clay, 1992).<br />

The Over-Exploitation <strong>of</strong> Forest Products<br />

Expanding production to enter new markets, or fill increased demand in existing markets, can result<br />

in production problems. There exists a lag time between increased demand and adequate supplies,<br />

as producers attempt to develop systems that will supply the quantity and quality <strong>of</strong> the product<br />

desired (FAO, 1991). The harvesting <strong>of</strong> forest products for subsistence or small-scale trade locally<br />

is usually based on traditional systems <strong>of</strong> forest management which have, through trial and error over<br />

many years, developed sustainable methods <strong>of</strong>extraction. When demand increases, however, existing<br />

harvesting methods usually will not suffice; additionally, collectors from outside the area will enter<br />

the forest to harvest valuable products 12 (Padoch, 1992; Cunningham, 1991; Vasquez and Gentry,<br />

1989; Falconer, 1990; Godoy et al., 1992). Vasquez and Gentry (1989) describe this as the "critical<br />

behavioral bottleneck" in the process <strong>of</strong> conversion from local consumption to commercial<br />

exploitation.<br />

Traditional resource management usually involves inadvertent (taboos, religious, seasonal and social<br />

restrictions on gathering and the nature <strong>of</strong> equipment used) and at times direct management <strong>of</strong><br />

resources if harvesting practices appear to create shortages in a resource <strong>of</strong> value to society or the<br />

socio-political nature <strong>of</strong> society depends upon it (Cunningham, 1991; Davies and Richards, 1991;<br />

lain, 1992). When large demand for a forest resource is created, harvesting practices <strong>of</strong>ten change<br />

in communities that have entered the cash economy, suffer from increasing unemployment, or have<br />

undergone cultural change 13 (Cunningham, 1991; Vasquez and Gentry, 1989). Protecting these<br />

resources by establishing laws or policing them will not, however, protect them from exploitation (De<br />

long and Mendelssohn, 1992; Peluso, 1992).<br />

Transportation<br />

The costs <strong>of</strong> transportation and the perishability <strong>of</strong> a forest product determines how far away products<br />

will be harvested 14 (De long and Mendelssohn, 1992). NTFPs tend to be difficult to ship due to<br />

12 For example, pau rosa (Aniba roseadora) is currently over-harvested to the point <strong>of</strong> elimination in the Amazon,<br />

but could 'be replanted and sustainably managed. Copaiba oil (Copaifera multijuga) can easily be harvested by drilling<br />

a hole with a brace and bit and tamping it, but is more <strong>of</strong>ten destructively harvested by cutting a hole with an axe.<br />

13 For example, Wamulwange (1993) reports from Zambia: "In the past the traditional government would charge a<br />

sum <strong>of</strong> five pounds for cutting timber without a license. One pound for cutting a fruit tree, but no charge for the<br />

collection <strong>of</strong> fIrewood. Some areas were declared as forest reserves where a culprit would be charged the sum <strong>of</strong> ten<br />

pounds for grazing, ten pounds for cutting and one pound for trespassing. People are not respecting these laws anymore."<br />

14 Assai is found in nearly monocrop stands, both naturally occurring and created, near the mouth <strong>of</strong>the Amazon and<br />

in varying densities along the river all the way to Bolivia. In 1990, the market for assai fruit, which was entirely domestic,<br />

was nearly US $100,000,000. The main market for the fruit is in Belem, where as many as 50,000 I <strong>of</strong> unprocessed fruit<br />

are sold daily. Harvesters in the area earn as much as US$10 to $US15 per hour. The Belem market can only be supplied<br />

by fruit that is no further than a single night's boat trip away because after 24 hours it spoils. Within this area, then, assai<br />

is very valuable. Outside <strong>of</strong> this radius, assai stands are decimated and sold for palm heart (Bovi and de Castro in: Clay<br />

and Clement, 1993; Clay and Clement, 1993).


spoilage, although many have a much higher kilogram value than even the most expensive timber<br />

(Clay and Clement, 1993; Richards, 1993).<br />

Box 1: Lesser Known Timber Species<br />

In addition to non-timber forest products, the use <strong>of</strong> a larger number <strong>of</strong> timber species in a natural<br />

forest can increase yield/unit area, and could improve the economic outlook <strong>of</strong> natural forest<br />

management. Generally, only 10-50 <strong>of</strong> the thousands <strong>of</strong> tropical tree species have been commercially<br />

exploited, but high-grading has depleted stocks <strong>of</strong> premium timbers and deforestation has made timber<br />

scarce (Hartshorn, 1990). As a result, efforts have been undertaken to promote the lesser known species<br />

(LKS) <strong>of</strong> high-diversity tropical forests. Although if done to excess, this could result in the depletion <strong>of</strong><br />

nutrients in forest ecosystems, it is generally accepted that - if done sustainably - the harvest <strong>of</strong> LKS<br />

would make better use <strong>of</strong> the forest resource, and less forest could be harvested to meet demands for<br />

timber. In order for LKS to succeed, the conservatism within the timber trade must be overcome,<br />

species must be marketed 10 consumers, infonnation on the wood properties <strong>of</strong> these species must be<br />

made available, and consistent supplies ensured (Jonson and Lindgren, 1990).<br />

In general, the more distant a timber consumer market is from its timber supply area, the more selective<br />

it is about the number <strong>of</strong> wood species it consumes. In part, this is due to the costs <strong>of</strong> transJX>rt. Local<br />

markets in the Amazon, for example, appear to be the least discriminating, with national markets<br />

accepting a large percentage less and international markets even fewer (Browder, 1986; Feamside,<br />

1990). Browder found, for example, that <strong>of</strong> the 7 principle Brazilian Amazon woods exported to the<br />

United States in 1982, mahogany accounted for 84%.<br />

In Queensland only one species, Red cedar (Toona Australis) was used by the early settlers, 10 species<br />

were used by 1930, and over 100 species by 1945. Leslie (1987) suggests that this is a function <strong>of</strong> the<br />

growing scarcity <strong>of</strong> prime species and thus the greater acceptance <strong>of</strong> alternatives and <strong>of</strong> small-sized<br />

species.<br />

In Malaysia, only 30 timber species are harvested for export in significant quantities (JoOOs, A.D.,<br />

1992). In the Brazilian Amazon, Uhl et al. (1989) reported that 222 trees <strong>of</strong> only 30 species were<br />

extracted from a 52 ha forest tract in Para, although the total number <strong>of</strong> timber species in the Amazon<br />

is estimated by Martini et al. (1993) at 335, and local mills now use approximately 140 species (as<br />

opposed to 6 species 40 years ago). Nair (1991) found that for an area in India, 30 species were listed<br />

for felling, but in fact a disproportionately large number belonging to 3 or 4 species were removed for<br />

specific ends.<br />

Martins (1992) reports that the promotion <strong>of</strong> lesser-known or non-traditional wood species in Honduras<br />

looked promising. Because the traditional mahogany and Spanish cedar species are increasingly scarce,<br />

room in the market has been created for the LKS species. Promotional activities, such as an exhibit <strong>of</strong><br />

fInished furniture pieces <strong>of</strong> high quality made entirely <strong>of</strong> LKS, have proved to be a success (Martins,<br />

CIDA, pers. comm., 1992).<br />

9


10<br />

Box 2: The Economic Valuation <strong>of</strong> Non-Timber Forest Products<br />

The harvest <strong>of</strong> medicinal plants for local use in Belize was estimated to have a net present value (NPV)<br />

<strong>of</strong> $726/ha in 30 year old forest and $3,327/ha in 50 year old forest. The plots in which medicinals<br />

were collected also contained other valuable income-generating NTFPs, like chicIe (Manilkara zapota)<br />

and allspice (Pimenta dioica), which were not included in these values. This compares favourably with<br />

alternative land uses. Estimates <strong>of</strong> the value <strong>of</strong> intensive agriculture in the Brazilian rainforest are<br />

$339/ha, and in Guatemalan forests $288/ha. The most successful pine plantations proposed for the<br />

tropics yield only $3,184/ha (Balick and Mendelssohn, 1992).<br />

Peters et al. (1992) similarly demonstrated the economic superiority <strong>of</strong> NTFPs using standard costbenefit<br />

analysis for a forest in Peru. The NPV <strong>of</strong> sustainable fruit and latex harvests was estimated at<br />

$6,330/ha, and <strong>of</strong> total NTFPs at $6,820. The standing volume <strong>of</strong> merchantable timber in this tract <strong>of</strong><br />

forest was 93.8 m 3 jha if liquidated in one felling, and a net revenue <strong>of</strong> $1,000 would result upon<br />

delivery to the sawmill. A logging operation <strong>of</strong> this intensity, however, would so damage the residual<br />

stand that future revenues from the site would drastically reduce or eliminate future harvests <strong>of</strong> NTFPs.<br />

The net financial gains from timber extraction would be reduced to zero if as few as 18 fruit trees were<br />

damaged by logging. Periodic selective cuttings <strong>of</strong> the 60 commercial species at the site, however,<br />

would produce volumes <strong>of</strong> 30 m 3 /ha every twenty years, which would produce net revenues <strong>of</strong> about<br />

$310 at each cutting cycle, or an NPV <strong>of</strong> $490 (peters et al., 1989).<br />

In Ecuador, Grimes et al. (1993) found that the three plots in their study produced NPVs for NTFPs <strong>of</strong><br />

$2939/ha, $2721/ha, and $1257/ha (with upland forest providing the largest returns). Timber resources<br />

calculated on a 40 year rotation length produced an NPV <strong>of</strong> $188/ha. Cattle ranching has a NPV <strong>of</strong> $57<br />

in the same area. Markets for NTFPs appeared to be expanding locally, regionally, and nationally. The<br />

values calculated for NTFPs do not include medicinal herbs or shrubs, flowers, wildlife, tourism or the<br />

wide range <strong>of</strong> environmental services provided by intact forests, all <strong>of</strong> which would greatly enhance the<br />

forest's economic value. For example, one guanta (Paca agouti), a forest rodent, sells for $20.69 in the<br />

local market, and decorative butterflies sell to tourists for a similar price. Selective, low-impact<br />

harvesting <strong>of</strong> timber species would additionally add to the NPV for the forest area (Grimes et al.,<br />

1993).<br />

The results <strong>of</strong> these valuations <strong>of</strong> non-timber products will vary depending upon the biological and<br />

economic diversity <strong>of</strong> study sites, temporal changes in market prices, production levels and harvest<br />

intensities, and the methods and reporting procedure employed by the researchers15. Although these<br />

studies usefully demonstrate the previously ignored "invisible" economic value <strong>of</strong> NTFPs, the<br />

methodology employed is still in development and there remain a number <strong>of</strong> doubts as to the practical<br />

relevancy <strong>of</strong> the results. For example, many <strong>of</strong> these valuations do not take into consideration the<br />

impact <strong>of</strong> greatly increased supplies on the market and prices, transport costs from more remote areas,<br />

the non-sustainable nature <strong>of</strong> some harvesting, the shortcomings <strong>of</strong> basing values on returns to land<br />

rather than labour (which is the scarce factor), the assumption that local populations would base<br />

decisions on discounted future values, and the great importance <strong>of</strong> institutional factors (such as insecure<br />

land tenure) that influence local decisions about the use <strong>of</strong> forest resources (Amold, pers. comm., 1994;<br />

Godoy and Lubowski, 1992; Pinedo-Vasquez et al., 1992).<br />

IS Godoy and Lubowski (1992) suggest that because many <strong>of</strong> these studies measure costs, quantities extracted, and<br />

prices in different ways, the results cannot be directly compared with each other; researchers, therefore, must pay more<br />

attention to their methods so that future studies can produce genemlizable results (Godoy and Lubowski, 1992).


12<br />

Because <strong>of</strong> the unpredictable nature <strong>of</strong> markets for non-timber products and the irregularity <strong>of</strong> yields<br />

from these species, it is generally best to manage forests for a wide range <strong>of</strong> species and potential<br />

products 18 •<br />

When selecting species for natural forest management, Peters (1993) suggests that, given a group <strong>of</strong><br />

species with similar economic pr<strong>of</strong>iles, ecological criteria should be incorporated to assess the<br />

potential for sustained-yield management: the life cycle characteristics, the multiplicity <strong>of</strong> uses and<br />

types <strong>of</strong> resources produced, the abundance in different forest types, and the size-class distribution<br />

<strong>of</strong> populations.<br />

Species that are annually fruiting, hermaphroditic, and pollinated by a common, generalist vector are<br />

much easier to work with than an unpredictably fruiting, dioecious species requiring specialized<br />

animal vectors for pollination and seed dispersal 19 • Primary forest species adapted for growth and<br />

regeneration under a closed canopy, such as Carapa guianensis, Copaijera multijuga, Theobroma<br />

grandijlorum, and Coumarouna odorata in the Amazon will, in most cases, be preferable to fastgrowing<br />

early pioneers which require large gaps for seedling establishment (Peters, 1993; Smith et<br />

al., 1992). Following logging, however, larger-sized gaps can be filled with useful pioneer species,<br />

such as Croton cajucara in the Amazon or Rauvolfia vomitoria and Musanga species in West<br />

Africa 20 •<br />

The abundance, or current density and distribution, <strong>of</strong> a species will also affect the ease with which<br />

they can be managed, their yield per unit area, and the likelihood that they can be sustainably<br />

managed (Peters, 1993; Gentry, 1992; Reining and Heinzman, 1992; Reining et al., 1992;<br />

18 For example, a good year <strong>of</strong> Brazil nut harvest is usually followed by a bad one, as the tree uses most <strong>of</strong> its<br />

accumulated reserves and takes more than a year to accumulate more. Fruiting is said to be heavy in alternate years only<br />

(Clay and Clement, 1993). The quantity <strong>of</strong> copaiba (Copaijera multijuga) drawn from each tree depends upon the<br />

botanical species <strong>of</strong> the tree, its age, the lapse <strong>of</strong> time since the previous tapping and upon season.<br />

19 The Brazil nut (Bertholletia excelsa), for example, has well-established and lucrative markets, but is cross pollinated<br />

by non-social or semi-social large bees with enough strength to pry open flowers, and has seeds which are dispersed by<br />

agoutis (Dasyprocta aguti) and squirrels (Sciureus spp.), which appear to be the only species capable <strong>of</strong> gnawing through<br />

the extremely woody pericarp (Mori, 1992).<br />

A lack <strong>of</strong> Brazil nut regeneration on the forest floor has been observed throughout Amazonian, but has not been<br />

adequately explained. It is possible that, because the large capsule fruits are so easy to find, gatherers do not leave enough<br />

seed on the forest floor to be dispersed by agoutis and develop into seedlings. In addition, lack <strong>of</strong> regeneration could be<br />

due to reduced populations <strong>of</strong> the major dispersal agent (agoutis) through hunting, low fertilization/germination rates, seed<br />

vulnerability to fungus, seedling dependence on large openings, the burning beneath adult trees to facilitate collection and<br />

nut production, or a lack <strong>of</strong> pollinators in cases where trees are not in proximity to the natural forest on which the bees<br />

depend (Nepstad et al., 1993; Mori, 1992; Smith et al., 1992).<br />

20 Rauvolfia species have yielded reserpine, the starting point for traditional and phannaceutical treaunents for<br />

hypertension. The crushed root, seeds and leaves are used locally in a variety <strong>of</strong> traditional remedies. The ashes <strong>of</strong>burned<br />

Musanga spp. are used as a salt substitute, its wood for a variety <strong>of</strong> construction purposes, and various plant parts are<br />

used in traditional medicines (Abbiw, 1990).


Cunningham, 1992)21. Clearly, rare species are in more danger from the essentially random damage<br />

inflicted by logging operations than common species are (lohns, A.D., 1992); similarly, unsustainable<br />

harvesting <strong>of</strong> plant parts or entire trees for non-timber forest products will endanger populations <strong>of</strong><br />

rare individuals more than the common. Although high-density populations <strong>of</strong> species are easier to<br />

manage than low, the vast majority <strong>of</strong> tropical species occur at low densities (Sayer, 1991; Peters,<br />

1993; Peters et al., 1989). Additionally, most traditional management systems promote and depend<br />

upon a wide-range <strong>of</strong> forest vegetational zones and products, and so are directly linked to high<br />

species diversity. (Nair, 1991; Salick, 1992; Toledo, 1992; Alcorn, 1990; Phillips, 1989)22. In<br />

economic tenns, the harvest <strong>of</strong> a multiplicity <strong>of</strong> non-timber forest products is preferable to the harvest<br />

<strong>of</strong> a few because it minimizes the potential impacts <strong>of</strong> individual product price fluctuations (Grimes<br />

et al., 1993).<br />

In the Amazon, most <strong>of</strong> the major non-timber product species occur at extremely low densities, but<br />

have wide geographic distributions. Carapa guianensis, Hymenaea courbaril, Platonia insignis,<br />

Dipteryx odorata, and Caryocar villosum average less than one tree!hectare (although densities <strong>of</strong>4-6<br />

trees!ha have been reported for Carapa in northeast Costa Rica). Copaijera multijuga averages just<br />

over 1 tree/ha (Clay and Clement, 1993; lonson and Lindgren, 1990). Theobroma grandiflorum can<br />

reach 14 trees!ha and Bertholletia excelsa as many as 20 trees!ha, but this type <strong>of</strong> abundance results<br />

most likely from Amerindian intervention 23 •<br />

Many timber species also provide non-timber products, which can generate revenue throughout the<br />

economically "unproductive" building phase following logging operations (Salick, 1992; Reining and<br />

Heinzman, 1992; Malhotra et al., 1991)24. Martini et al. (1989) found that <strong>of</strong> the 335 timber species<br />

utilized in the Brazilian Amazon, one-third were also valued for their fruits, medicinal properties,<br />

and/or gums and resins. The bulk <strong>of</strong> these have "high" (but not greater than $40/m 3 ) wood values and<br />

are under "high" logging pressure (Martini et aI., 1993). In some cases, such as rosewood (Aniba<br />

duckei), non-timber values exceed timber values (Clay and Clement, 1993). The physiological trade<strong>of</strong>fs<br />

<strong>of</strong> harvesting multiple resources from a single tree should be assessed prior to initiating such a<br />

21 In the Maya Biosphere Reserve in Peten Guatemala, three important NlFPs occur in remarkably high densities:<br />

xate (Chamaedorea spp.), the leaves <strong>of</strong> which are exported to the United States and Europe, where they serve as a green<br />

backdrop for cut flower arrangements; chicle (Manilkara zapota), a natural latex used in the manufacture <strong>of</strong> chewing gum<br />

bases; and allspice, the berry <strong>of</strong> Pimienta dioica used as a condiment, to preserve fish, and as a flavouring and curing<br />

agent in processed meats and bakery products. It is likely that this results from Mayan forest gardens and management.<br />

As Reining and Heinzman (1992) described it: "it is as if this is a pre-managed forest". High density and numbers <strong>of</strong><br />

these NlFPs and dozens <strong>of</strong> secondary timbers, thatching palms, construction materials and medicinal plants facilitate<br />

sustainable management in this area (Reining and Heinzman, 1992; Reining et al., 1992).<br />

22 An ethnobotanical study <strong>of</strong> eight indigenous groups in the tropical moist zones <strong>of</strong> Mexico showed that 1,380 plant<br />

species were identified as having one or more uses. Of these, 465 are primary forest species, 712 are secondary forest<br />

species (153 <strong>of</strong> the two preceding figures are species found in both), and 356 are species without ecological specification<br />

<strong>of</strong> habitat (Toledo, 1992).<br />

23 Forests dominated by stands <strong>of</strong> Brazil nut trees are called castanhais. They occur over an area <strong>of</strong> 8000 km 2 in the<br />

Amazon Basin (Balee, 1989).<br />

1A For example, in Southwest Bengal the pioneer species Diospyrros melnoxylon is harvested to provide a wrapper<br />

for Indian cheroots (bedi) in degraded forest prior 10 shading out by the canopy <strong>of</strong> timber species (Shorea robusta).<br />

Enrichment planting <strong>of</strong> mushrooms and medicinal also contributed significantly to economic returns after canopy closure<br />

(Malhorta, et al., 1991).<br />

13


14<br />

program; for example, tapping oleo-resin or latex will undoubtedly have some effect on the<br />

production <strong>of</strong> fruit (Peters, 1993).<br />

For the vast majority <strong>of</strong> non-timber forest species, we have little data on tree density, productivity<br />

and population structure. Many species are as yet undefined 25 • In order for sustained-yield <strong>of</strong> these<br />

species to be possible, however, we must know: the location and identity <strong>of</strong> the resource; the<br />

resource's history <strong>of</strong> exploitation; the total number <strong>of</strong> harvestable trees per hectare; the current<br />

population structure or size-class distribution <strong>of</strong> the species and the quality <strong>of</strong> regeneration; the<br />

productive capacity <strong>of</strong> the species being exploited, and the relationship between tree size, site, and<br />

productivity; and the maximum amount that can be exploited - i.e. the sustainable harvesr 6 •<br />

Plant Parts Used In Non-Timber Forest Products<br />

The sustainable harvesting <strong>of</strong> non-timber forest products depends in large part on the part <strong>of</strong> the plant<br />

that is used (Cunningham, 1991; 1992). The bulk <strong>of</strong> NTFPs can be grouped into three basic<br />

categories based on the type <strong>of</strong> plant tissue or compound actually exploited: reproductive propagules<br />

(fruit, nut/seed, oilseed); plant exudates (latex, resin, and floral nectar); and vegetative structures<br />

(stem, leaf, root bark, and apical bud) (Peters, 1993).<br />

Reproductive Propagules<br />

The harvest <strong>of</strong> fruits, nuts and oilseeds removes plant embryos from the site, and hence reduces the<br />

total number <strong>of</strong> seedlings that can potentially be recruited into the tree population under exploitation,<br />

alters populations structure, genetic variability, and can impact the foraging behaviour <strong>of</strong> local animal<br />

populations 27 • Low density populations which display an obligate relationship with a specific<br />

pollinator or seed disperser will have a much lower sustainable yield <strong>of</strong> fruit, and will be more prone<br />

25 For example, the most important construction timber <strong>of</strong> the Iquitos, Peru area (72-74% <strong>of</strong> the wood used for general<br />

construction came from this species) was recently collected by botanists and found to be a species new to science. "Boa<br />

caspill", the main construction timber at Jenaro Herrera on the Rio Ucayali, turned out to be an undescribed species <strong>of</strong><br />

Havloclathra, so distinctive it was frrst suspected to be a new genus (Gentry, 1992).<br />

26 The most immediate and cost-effective way <strong>of</strong> doing this is by monitoring (in permanent sample plots) the<br />

population impact <strong>of</strong> exploitation and sequentially adjusting harvesting levels over time to obtain a sustainable yield; this<br />

method is known as successive approximation. In practice, achieving sustained yield through successive approximation<br />

will probably involve a considerable number <strong>of</strong> harvest adjustments. There is frequently a tag time in the demographic<br />

response <strong>of</strong>a population to perturbations, such as logging or high-intensity harvest, and after several cycles <strong>of</strong> apparently<br />

stable results from PSPs, populations may exhibit drastic fluctuations in seedling and sapling densities. Sustainable<br />

harvesting levels can thus be achieved through a process <strong>of</strong> reciprocal feedback and "fine-tuning" (peters, 1993; Malhorta<br />

et al., 1991). Frequently, too, local peoples have determined sustainable levels <strong>of</strong> harvest from trial and error, and<br />

ethnobotanical information can reveal a substantial amount <strong>of</strong> information concerning the sustainable productivity <strong>of</strong><br />

tropical forest resources (De long and Mendelssohn, 1992).<br />

27 If a tree population produces 1,000 seeds and 95% <strong>of</strong> the new seedlings produced from these seeds die during the<br />

frrst year, the population still has recruited 50 new progeny. If, on the other hand, intensive fruit harvesting removes all<br />

but 100 <strong>of</strong> these seeds from the site prior to germination, the maximum number <strong>of</strong> seedlings that can be recruited into<br />

the population is reduced to only 5. This ten-fold shortfall in recruitment rate can cause serious changes in the structure<br />

<strong>of</strong> the population (peters, 1993).


to over-exploitation than species with abundant regeneration, not prone to extreme post-dispersal seed<br />

predation, that are pollinated and dispersed by either "generalist" animals or abiotic factors (Peters,<br />

1993)28.<br />

In some cases, poor harvesting practices can lead to drastic species decline not associated directly<br />

with removal <strong>of</strong> the reproductive propagules. For example, trees are <strong>of</strong>ten felled to collect fruits. In<br />

the Peruvian Amazon, female trees <strong>of</strong> the dioecious aguaje palm (Mauritia fleusosa) are felled by<br />

commercial collectors and after a few such "harvests" the forest is left with a preponderance <strong>of</strong><br />

barren male trees (Peters, 1993).<br />

Plant Exudate<br />

When properly conducted, the tapping <strong>of</strong> latex, resins and gums does not disturb the forest canopy,<br />

kill the exploited tree or remove its seed from the site (Peters, 1993). In theory, this activity can<br />

easily be sustainable, but in practice poor harvesting techniques plague these products, as well.<br />

Copaiba (Copaifera multijuga), for example, can be tapped through its lifespan if a brace and bit are<br />

used, and holes are filled following tapping. In practice, however, axes are <strong>of</strong>ten used to create holes<br />

and the wounds created do not heal (Clay and Clement, 1993). Additionally, many species, such as<br />

the Amazonian Couma macrocarpa, which produces valuable fruits and latex, are felled to harvest<br />

plant exudate that could be easily tapped 29 (Peters, 1993). It is also likely that repeated tapping, for<br />

example <strong>of</strong> Hevea brasiliensis, reduces vegetative growth and total fruit production (Peters, 1993).<br />

Vegetative Structures<br />

Although the origin and use <strong>of</strong> plant products from root, stems, leaves, barks, or apical buds, is very<br />

different, their harvest produces a similar ecological impact. The plant species will either be killed<br />

during harvest or will survive and regenerate the vegetative structures removed (Peters, 1993).<br />

Stems<br />

Rattan harvest in Southeast Asia is an example <strong>of</strong> particularly destructive harvesting <strong>of</strong> commercial<br />

quantities <strong>of</strong> stem fibre. Rattan is harvested by cutting the plant at the base and then pulling the<br />

(<strong>of</strong>ten more than 100 m long) stems and leaves out <strong>of</strong> the forest canopy. Large-caned, single<br />

stemmed rattan, such as Calamus manan, do not resprout after cutting, so harvesting kills these<br />

individuals. Small-caned, multi-stemmed individuals, such as C. caesius and C. trachycoIus, can<br />

eventually resprout after harvesting, and individual stems can be cut from the clumped vegetation<br />

every 2-3 years after an initial 8-10 years <strong>of</strong> growth (de Beer and McDermott, 1989; Peters, 1993).<br />

Over-exploitation <strong>of</strong> both small- and large-caned stems, including cutting them too close to the<br />

28 The illipe nut <strong>of</strong> Shorea species in Southeast Asia, for example, produces a high quality oil used in the cosmetic<br />

industry and as a substitute for cocoa to harden chocolate. But many <strong>of</strong> the most important Shorea species providing illipe<br />

nut are fruit masting, and set flowers every five years, so productivity is low (De Jong and Mendelssohn, 1992; Dixon<br />

et al., 1991).<br />

29 Gaharu wood (Aquilaria spp.) is an extremely valuable aromatic resin resulting from infections <strong>of</strong> trees by unknown<br />

pathogens. Aquilaria spp. are found in the primary forests <strong>of</strong> Southeast Asia. The collection <strong>of</strong> gaharu usually involves<br />

felling the trees, which resprout weakly, if at all. Because there are no external signs to indicate whether a tree contains<br />

the valuable exudate, collectors frequently fell every Aquilaria tree they find (although the Penan are successful at<br />

identifying the desirable trees without felling). Uncontrolled exploitation and wasteful felling in search <strong>of</strong> gaharu has led<br />

o its rapid decline (Padoch et al., 1992; Peters, 1993; Zemer, 1993; Stockdale, pers. comm., 1993).<br />

15


16<br />

ground and at too young an age, coupled with increased logging pressures on the forests in which<br />

the vast majority <strong>of</strong> commercial species are harvested, have led to increasing scarcity <strong>of</strong> valuable<br />

rattan species (Peters, 1993; Sayer, 1991; De Jong and Mendelssohn, 1992)30.<br />

The harvest <strong>of</strong> chewsticks, made from the stems <strong>of</strong> primarily Garcinia epunctata and Garcinia afzelii,<br />

provides the main means <strong>of</strong> dental care for 90% <strong>of</strong> the population in southern Ghana. Despite the<br />

increasing popularity <strong>of</strong> Western-style toothbrushes, toothpaste is hard to come by, and demand for<br />

chewsticks continues to rise (Falconer, 1992; Cunningham, 1992). Trees are cut from the forests,<br />

usually by groups <strong>of</strong> gatherers who come out from the cities specifically for this purpose, and the<br />

harvest is then taken back by truck to the urban markets. Falconer (1992) found that 2,000-6500 trees<br />

a month are harvested in the Kumasi region alone, and that almost every gatherer claims that these<br />

species are becoming increasingly scarce in the forest.<br />

Leaves<br />

The harvesting <strong>of</strong> leaf fibres may have a negligible effect on the structure and abundance <strong>of</strong> plant<br />

populations being exploited if: individual plants are not killed in the process, a few healthy leaves<br />

are left on each plant to photosynthesize, reproductive structures and meristems are not damaged, and<br />

sufficient time is allowed between successive harvests for the plant to produce new leaves (Peters,<br />

1993). Many traditional medicines and internationally-traded herbal medicines come from leaves 31 •<br />

In many cases, the extent <strong>of</strong> demand for leaves has resulted in diminished populations, due to the<br />

factors described above. Leaves stripped from the Amazonian Pilocarpus spp., for example, provide<br />

the active compound pilocarpine, which is used in the pharmaceutical treatment <strong>of</strong> glaucoma 32 • In<br />

1989, over 1300 tons <strong>of</strong> dried plant material were exported to the United States alone (Elisabetsky,<br />

1991). Due to scarcity and the desire to control supply, the pharmaceutical company E. Merck <strong>of</strong><br />

Germany has spent ten years attempting to cultivate Pilocarpus spp. It has had a difficult time doing<br />

so, as leaves with adequate quantities <strong>of</strong> the desired chemical compounds have been difficult to<br />

produce in plantation (Sheldon et al., 1993). The secondary compounds useful in medicines are <strong>of</strong>ten<br />

not produced when plants are grown outside <strong>of</strong> the difficult and competitive atmosphere <strong>of</strong> their<br />

natural environment. As a result, wild sources will continue to be an important source <strong>of</strong> raw material<br />

for even the most "high-tech" <strong>of</strong> medicines.<br />

Roots and Bark<br />

The collection <strong>of</strong> roots and bark usually kills or fatally weakens the exploited tree species,<br />

particularly with high-intensity harvests (Peters, 1993). Bark can be renewably harvested if trees are<br />

30 Stockdale (pers. comm., 1993) reports that traditional management and harvesting practices in East Kalimantan have<br />

<strong>of</strong>ten been overlooked by outside collectors and are being dismissed by some <strong>of</strong> the younger members <strong>of</strong> the community.<br />

For example, the traditional practice <strong>of</strong> cutting the stem 1 m from the ground and then bending the tip <strong>of</strong> the stump back<br />

into the earth to prevent fungal infection spreading from the stump into the rest <strong>of</strong> the clump, is perceived as wasteful<br />

<strong>of</strong> valuable cane.<br />

31 For example, sacaca (Croton cajuacara) leaves, dried and made into teas or sold in gelatin capsules, have become<br />

"the most" fashionable medicinal plant in Belem these days. It is used to treat diabetes, diarrhoea, and to lower<br />

cholesterol. Because it is a weedy pioneer species, its survival will probably not be severely affected even by the rapidly<br />

increased demand that has recently occurred (Venturieiri and Ribeiro in: Clay and Clement, 1993).<br />

32 In 1989, the estimated sales <strong>of</strong> pilocarpine in the United States alone reached US$28,000,000 (Elisabetsky, 1991).


not girdled, but management strategies <strong>of</strong> this kind are <strong>of</strong>ten disregarded, or succumb to enonnous<br />

market demand for the species 33 • The harvest <strong>of</strong> roots will kill the individual plant, but management<br />

<strong>of</strong> these species, including regulating <strong>of</strong>f-take, can minimize the impact on populations 34 • Both barks<br />

and roots can be harvested from multi-purpose trees felled for timber, prior to or post-harvesting. The<br />

bark <strong>of</strong> Pau d'arco (Tabebuia serratifolia), for example, is harvested for use in medicines used<br />

regionally and internationally, prior to logging operations in Brazil.<br />

Apical Buds<br />

Palm hearts are the most important and well-known example <strong>of</strong> apical bud use. In a single-stemmed<br />

palm species, harvesting the "heart", or apical meristem, necessarily kills the tree. Euterpe precatoria,<br />

for example, was widely harvested for a palm heart canning factory near Iquitos, Peru, which was<br />

subsequently forced to close due to scarcity <strong>of</strong> raw materials. The multi-stemmed growth fonn <strong>of</strong> E.<br />

oleracea in the Amazon enables the species to sprout back after cutting, and collectors in the eastern<br />

Amazon have taken advantage <strong>of</strong> this to manage for a sustained-yield harvest (Peters, 1993).<br />

Wildlife as a Non-Timber Forest Product<br />

Wild game is <strong>of</strong>ten overlooked as an important non-timber forest product, but is one <strong>of</strong> the most<br />

marketable and higher value forest products. The rural population's dependence on wild meat for<br />

subsistence is also extremely important. For the hunter and his family, hunting is free or very cheap<br />

in monetary tenns. This is important because large sectors <strong>of</strong> the rural population still have limited<br />

access to cash with which to buy substitutes (Caldecott, 1988).<br />

Caldecott (1988) calculated that about half the minimum protein supply in Sarawak is derived from<br />

fresh meat, almost 60% from the wild, including fish. Meat also provides a "buffer" in times when<br />

other crops are inadequate 35 • Hunting constitutes the major source <strong>of</strong> yearly revenue for the families<br />

in the northern Congo studied by Wilkie et al., (1991). Many families reported that without revenue<br />

obtained from hunting, they would be unable to buy cooking utensils, clothing, medicine, and<br />

educational materials for children. Logging operations in this area had directly and indirectly<br />

contributed to the depletion <strong>of</strong> local stocks <strong>of</strong> wildlife (Wilkie et al., 1991).<br />

33 The bark from the Cameroonian species Pygeum ajricanus, for example, is used to treat urinary problems associated<br />

with enlarged prostate glands. By removing bark from opposite quarters every 5 years, and stripping up to 20 m above<br />

ground, the bark <strong>of</strong> P. africanus can probably be sustainably harvested, as it appears to be very resilient to bark removal.<br />

In practice, however, trees are <strong>of</strong>ten felled to collect the bark, or entire trees are stripped. By 1989, most <strong>of</strong> the forests<br />

where the bark was sOUTced in Cameroon were depleted <strong>of</strong> P. africanus (Cunningham, 1993; Sheldon et al., 1993).<br />

34 The root <strong>of</strong> the slow growing understorey herb ipecac (Cephaelis ipecacuanha) has long been used traditionally<br />

to treat dysentery and as an expectorant. The scale <strong>of</strong> current international demand for ipecac (as an ingredient in<br />

expectorants and in a treatment for amoebic dysentery) has led to over-harvesting in the wild and severe reductions in<br />

local populations (Sheldon et al., 1993; Husain, 1992). In Nicaragua, local communities successfully managed the<br />

extraction <strong>of</strong> ipecac, which is one <strong>of</strong> the most important NTFPs in the region, by planting beds below the forest<br />

overstorey (Salick, 1992).<br />

35 The contribution <strong>of</strong> wild meat to human nutrition in the interior is illustrated by rations for pupils at boarding<br />

schools: 203 tonnes <strong>of</strong> meat and fish were consumed at 63 schools in 1984-85; the largest single component was wild<br />

pig meat at 32%, with other wild meat contributing about 7%, fish about 18% and domestic pork, beef and chicken 13­<br />

16% each. In 1984 the estimated traffic in wild pig and deer meat exceeded $4,000,000 in the Rajang basin.<br />

17


18<br />

Wild game is also one <strong>of</strong> the most popular foods in Ghana. It is in equally high demand in urban and<br />

rural areas. Forests and fallow fields provide the habitat for many commonly consumed wildlife<br />

species. Forest-dependent animals are particularly important in humid West Africa because substitutes<br />

are <strong>of</strong>ten not available. Tsetse flies are endemic to this area and make cattle production difficult.<br />

Wild animals contribute between 20-90% <strong>of</strong> the total animal protein consumed (Falconer, 1992;<br />

1990). A.G. Davies (1987) similarly found that in the Gola Forest Reserve in Sierra Leone,<br />

unfavourable conditions for cattle and other livestock had lead to a traditional reliance on fish and<br />

bushmeat as sources <strong>of</strong> animal protein. While different communities have different preferences, for<br />

example Moslems disdain monkey meat, hunting <strong>of</strong> wildlife in this area is a major activity.<br />

In addition to the economic and nutritional importance <strong>of</strong> wildlife must be added the role <strong>of</strong> hunting<br />

in many indigenous cultures. Hunting is central to the culture <strong>of</strong> the Penan, many <strong>of</strong> whom are still<br />

hunter gatherers. It is also integral to shifting cultivation and its associated cultures. As Caldecott<br />

(1988) said, for centuries, "to be a Dayak was to be a shifting-cultivator, was to be a hunter".<br />

Redford (1991) notes that in the Neotropics many indigenous communities associate hunting with<br />

social status and power. In the sharing <strong>of</strong> game with the rest <strong>of</strong> the community, the hunter builds<br />

debts, acquires allegiances and contributes to social cohesiveness. Shortages in meat have been known<br />

to create social tension and even village fission, and hunters in some cases must engage in wage<br />

labour to purchase canned meat and fresh fish, with repercussions in the community cohesiveness.<br />

Decreases in game yields or prohibitions on hunting can have repercussions well beyond the strictly<br />

nutritional (Redford, 1991). (See below for a discussion <strong>of</strong> the "Impact <strong>of</strong> Natural Forest Management<br />

on Wildlife").<br />

v. The Ecological Underpinnings <strong>of</strong> Natural Forest Management<br />

The sustainable harvest <strong>of</strong> timbers from tropical forests is dependent upon a variety <strong>of</strong> factors and<br />

complex ecological relationships, most <strong>of</strong> which remain largely unknown. Despite ignorance <strong>of</strong> the<br />

intricacies <strong>of</strong> tropical forest ecosystems, general ecological principles have and continue to guide the<br />

design <strong>of</strong> natural forest management objectives and the harvesting operations and silvicultural systems<br />

they employ. These, in turn, influence the diversity <strong>of</strong> species and forest structure, and the harvest<br />

<strong>of</strong> non-timber forest products that rely on this diversity.<br />

Forest management - whether manifested in logging operations or silvicultural treatments - creates<br />

disturbances. It is in the extent that these disturbances mimic the size, duration and frequency <strong>of</strong><br />

naturally-occurring disturbance that ecological sustainability is achieved. Tropical forests are dynamic<br />

- not homogenous and static - mosaics <strong>of</strong> young, middle-aged and mature patches in different stages<br />

<strong>of</strong> regeneration (Putz and Brokaw, 1989; Uhl et al., 1990; Swaine, 1986)36. The concept <strong>of</strong> a steadystate<br />

"climax" forest has been questioned by many who feel that it does not reflect the dynamic<br />

nature <strong>of</strong> tropical moist forests, and some suggest that in fact forest virginity is "relative" (Hartshorn,<br />

36 Whitmore defmes three growth cycle phases in tropical forests: the gap phase, which is the opening <strong>of</strong> the canopy;<br />

the building phase, which consists <strong>of</strong> young trees, mostly shade-intolerant, growing rapidly to fIll the gap and attain the<br />

canopy; and the mature phase, formed by an intact canopy <strong>of</strong> large trees (Whitmore, 1991; Hartshorn, 1980).


1980; Niering, 1987)37. Not only have indigenous inhabitants long modified the forest environment,<br />

but small-scale natural occurrences such as treefalls, and large-scale disasters such as hurricanes,<br />

fires, landslides, and droughts, have created forests in a constant state <strong>of</strong> repair. Although a<br />

directional change in floristics will not occur in a mature phase forest as a whole, the composition<br />

at a given spot, from one tree to the next, will change. (Balee, 1989; Hartshorn, 1980; Whitmore,<br />

1991; 1992; Brown, 1992). Hartshorn (1980) found that in mature La Selva forest in Costa Rica, the<br />

turnover rate between two successive gaps occurring at the same spot in the forest was as high as<br />

118 +/- 27 years. Uhl et al. (1990) found that 4-6% <strong>of</strong> the forest area in Amazonian Venezuela was<br />

in a "gapped" condition at anyone time.<br />

Fundamental to most harvesting operations and silvicultural systems is the importance <strong>of</strong>regeneration<br />

in the gaps created by logging, and forest management systems have been developed to explicitly<br />

stimulate the growth <strong>of</strong> the pool <strong>of</strong> desirable species 38 . Gap size is manipulated during logging<br />

operations to match the growth characteristics <strong>of</strong> commercial species and silvicultural treatments are<br />

applied to release the advanced regeneration <strong>of</strong> desirable individuals and suppress the<br />

"undesirable"39 (Uhl et al., 1990; John, A.D., 1992, Wyatt-Smith, 1987). "Sustainable" forest<br />

management generally tries to mimic natural disturbances, and utilizes ecological succession and the<br />

mechanisms and factors that drive the processes <strong>of</strong> species change, population change and<br />

replacement through time to promote the regeneration <strong>of</strong> commercial species (Gomez-Pompa and<br />

Burley, 1991; Brown, 1992; Lawton, 1984). Interspecific competition for the establishment <strong>of</strong> sites<br />

has resulted in adaptive compromises in regenerative strategies, which results in varying competitive<br />

success <strong>of</strong> species in gaps. Differential species response will depend on the size <strong>of</strong> gaps, the timing<br />

<strong>of</strong> gap occurrence, the proximity and dispersal <strong>of</strong> seeds, growth rates, architectural construction that<br />

facilitates competition for light, nutrient requirements relative to competitors, nutrients available at<br />

the sight, water use and efficiency, and resistance to seed and seedling predation, pathogens and the<br />

increased herbivory associated with gaps (Denslow, 1980; Wyatt-Smith, 1987; Bazzaz, 1991; Nepsted<br />

et al., 1991; Hartshorn, 1980; Brown, 1992).<br />

37 Niering (1987) quotes Egler's 1947 statement that "the Climax and God have certain things in common for certain<br />

botanical atheists. To paraphrase Julian Huxley, the writer does not believe in the climax, because he thinks the idea has<br />

ceased to be a useful hypothesis. 1t<br />

38 The process <strong>of</strong> topling a tree, the fallen tree itself, together with the resultant hole in the canopy and accumulated<br />

debris on the forest floor is known as chablis or gap formation (Hendrison, 1990).<br />

39 The immigration <strong>of</strong> species and progress <strong>of</strong> succession can occur through a number <strong>of</strong> regeneration pathways:<br />

1) Seedlings and saplings <strong>of</strong> climax species present on the forest floor with little or no real growth toward adult size<br />

are stimulated by the new microenvironment resulting from canopy opening. UhI et al. (1990) found that in the<br />

Amazon, this advanced regeneration played a dominant role in tree fall gap formation and that four years after gap<br />

formation, composed 950/0 <strong>of</strong> the species> 1 m tall (Uhl et al., 1990; Bazzaz, 1991; Gomez-Pompa and Burley,<br />

1991).<br />

2) Sprouting from stem bases or roots following the destruction <strong>of</strong> above-ground plant parts also plays a major role<br />

in gap-phase regeneration (UhI et al., 1990; Irvine, 1989). Putz and Brokaw (1989) found in Panama that sprouted<br />

trees contributed more significantly to canopy closure in small gaps than in large, and that they initially grow faster<br />

than seedlings <strong>of</strong> the same species.<br />

3) Recolonization by the germination <strong>of</strong> pioneer species seed buried in soil seed bank, but not growing as seedlings<br />

or adult trees in the immediate area (Gomez-Pompa and Burley, 1991; Uhl et al., 1990).<br />

4) The arrival <strong>of</strong> seeds, which might fall from surrounding trees, or are dispersed from distant trees by wind, birds,<br />

bats, rodents and other mammals (UhI et al., 1990).<br />

19


20<br />

VI. Harvesting Operations in Managed Natural Forest<br />

The manner in which logging operations are conducted directly influences the present and future<br />

harvest <strong>of</strong> both timber and non-timber products from the forest. The extent <strong>of</strong> logging damage<br />

associated with most operations in the tropics results not only in the drastic decline <strong>of</strong> local species<br />

diversity and forest structural diversity, but also unfavourable rates <strong>of</strong> basal area growth <strong>of</strong><br />

commercial species through the destruction <strong>of</strong> seedlings, adolescent trees and soil surface and<br />

drainage pattems 40 (Whitmore, 1991; John, A.D., 1992; Dykstra and Heinrich, 1992; Dawkins,<br />

1958). Destructive logging operations can also result in increased fire hazards, lowland flooding, and<br />

decreased marine and coral reef productivity.<br />

Nepstad et al. (1991) found in the eastern Amazon that to harvest 1.6% <strong>of</strong> the trees in a forest, 12%<br />

<strong>of</strong> the trees with a dbh > 10 cm lost their crowns, 11 % were uprooted and 3% suffered substantial<br />

bark scarring. Uhl (1989) cites similar figures for residual stands in eastern Amazonian: 11%<br />

uprooted by bulldozers for roads, 12% crushed in felling, and 3% with excessive bark removal. In<br />

their efforts to extract 52 m 3 /ha, or eight trees, logging operators destroyed 26% <strong>of</strong> those remaining.<br />

Canopy cover (a summation <strong>of</strong> road area, felled tree area and log storage area) was reduced by half<br />

(Uhl, 1989; Wilkie et al., 1992). In Malaysia, A.D. John (1988) found that for the removal <strong>of</strong> 30-50<br />

m 3 /ha, or 3.3% <strong>of</strong> the trees, total canopy cover was reduced 50%.<br />

Damage in selective harvesting systems is usually patchy due to varying population densities <strong>of</strong><br />

commercial species. This is particularly pronounced in the highly diverse neotropical and African<br />

forests. In Malaysian dipterocarp forests, dominated by Dryobalanops aromatica or Shorea curtisii<br />

species, 14-72 trees may be removed per ha. Amazon terre firme forests yield on average 3-5<br />

trees/ha, and African forests can yield as few as 1.1 commercial trees/ha (John, A.D.; Uhl, 1989).<br />

Logging has numerous and potentially severe impacts on the soil surface <strong>of</strong> forests. The removal <strong>of</strong><br />

humus and topsoil during logging operations can result in high rates <strong>of</strong> soil erosion and leaching,<br />

particularly on slopes, and can produce high depositions <strong>of</strong> soil and debris on valley floors and flat<br />

areas during the wet season (Uhl et al., 1981). Rutting and the disturbance <strong>of</strong> the soil surface can also<br />

significantly disrupt the seed bank, seedlings, and superficial feeding roots critical to regeneration<br />

(Whitmore, 1991; Hendrison, 1990). But perhaps the most significant factor in the growth <strong>of</strong> future<br />

stands <strong>of</strong> timber and non-timber species on logged-over sites is soil compaction. Soil compaction<br />

affects tree growth through reduced porosity resulting in decreased diffusion <strong>of</strong> gases, root<br />

penetration, and the flow <strong>of</strong> water. The type and degree <strong>of</strong> soil compaction depends upon soil<br />

characteristics such as texture, structure, field moisture content, and organic matter content, and on<br />

the intensity <strong>of</strong> machinery traffic, its gross vehicle weight, steering system, and tire and track type<br />

(Hendrison, 1990; Jonson and Lindgren, 1990). Soil recovery is a slow process. Hendrison (1990)<br />

found in Suriname that skid trails used eight years previously were still maximally impacted. At Jarl,<br />

in eastern Amazonian Brazil, the clearance <strong>of</strong> large forest areas for plantations by machine was<br />

abandoned because trees grew so poorly on compacted soils (Whitmore, 1991).<br />

40 Damage to the residual stand can take the form <strong>of</strong> direct damage, wounds which are vulnerable to pathogen attack,<br />

water stress, the loss <strong>of</strong> symbiotic mycorrhizal infections, invasion by climbers due to excessive canopy opening, and<br />

windthrow resulting from increased turbulence <strong>of</strong> the uneven canopy (John, A.D., 1992; UbI, 1989).


Although adequate information exists to implement efficient and ecologically-improved logging<br />

systems, a number <strong>of</strong> economic and social factors have precluded its implementation. These include:<br />

short-term objectives that result in cost minimization and pr<strong>of</strong>it maximization; government concession<br />

agreements, incentives and payment schemes that do not stimulate sustained yield management; pr<strong>of</strong>it<br />

margins which have been unreasonably high in the past and have concessionaires "spoiled" with<br />

regards to expected returns; and the poor dissemination <strong>of</strong> knowledge from research to those<br />

conducting logging operations (lonson and Lindgren, 1990; Dykstra and Heinrich, 1992). But<br />

successful natural forest management depends on well-managed logging operations - as de Graaf and<br />

Poels (1990) said, "the first priority is to domesticate the logger, whereupon it is possible to<br />

domesticate the forest".<br />

A sustainable logging operation will involve more than felling and extracting timber. Harvest<br />

planning, technical supervision and post-harvest assessments that reflect concern about the biological<br />

diversity, non-timber forest products and long-term health <strong>of</strong> an ecosystem are required components<br />

<strong>of</strong> what Dykstra and Heinrich (1992) call "harvesting", as opposed to "logging", operations.<br />

Harvesting operations should be based on silvicultural considerations, as the success <strong>of</strong> silvicultural<br />

systems will in large part depend upon the preceding harvest. Wyatt-Smith (1987) describes timber<br />

felling as the first major silvicultural operation <strong>of</strong> a natural regeneration system. The key elements<br />

<strong>of</strong> a harvesting system are: harvest planning, forest roads, felling operations, skidding and yarding<br />

operations, and post-harvest assessments (Dykstra and Heinrich, 1992; Jonson and Lindgren, 1990).<br />

Harvest Planning<br />

Harvest plans are based on inventories and the collection <strong>of</strong> information necessary to ensure<br />

ecological sustainability and to plan transportation within the forest in a way that minimizes damage<br />

to residual stands and the total area disturbed by roads, landings, skid trails and cableways. All<br />

commercial and potentially commercial timber trees, trees to be felled (with a minimum <strong>of</strong> 20 m<br />

between two marked trees), seed trees, climbers to be cut, advanced growth for retention, mature<br />

species <strong>of</strong> importance for wildlife, important species and areas for local non-timber forest produce<br />

and biodiversity conservation, and filter/buffer strips for watercourses must be designated in the<br />

logging plan 41 • This information will result from integrated inventories conducted prior to the<br />

planning process, and should undergo subsequent monitoring through permanent sample plots and<br />

field visits to ascertain that management objectives have been met (ITTO Draft Guidelines, 1992;<br />

Dykstra and Heinrich, 1992; John, A.D., 1992; Jonson and Lindgren, 1990; Nair, 1991; Wilkie et al.,<br />

1992).<br />

The type <strong>of</strong> equipment to be used in a logging operation must be specified in a plan, as must the<br />

timing <strong>of</strong> the operation. The timing <strong>of</strong> logging operations will take into consideration rainy seasons,<br />

seedfalls, and the reproductive cycles <strong>of</strong> animals or species <strong>of</strong> non-timber value; contingency plans<br />

41 In the Guidelines for the Selective Logging <strong>of</strong> Rainforest Areas in North Queensland State Forests and Timber<br />

Reserves, for catchment areas> 60-100 ha, 30 m buffer strips are the required minimum for watercourses 20 m or more;<br />

20 m buffer strips are required for watercourses between 10-20 m wide; and 10 m buffer strips for those up to 10 m wide.<br />

The ITTO, in its Draft Guidelinesfor Conserving Biological Diversity in Forests Managed/or Timber, recommends 20<br />

m wide buffer strips along streams less than 20 m wide, and 50 m buffer strips along larger streams, rivers and lakes.<br />

These buffer strips provide vital habitat for many species <strong>of</strong> wildlife.<br />

21


22<br />

should be established for severe storms and other extreme events 42 • A floristic shift in diversity can<br />

result if parent trees are removed before fruiting or shedding <strong>of</strong> seeds occurs (Uhl et al., 1981). In<br />

the Malaysian Uniform System, the rule <strong>of</strong> thumb is that "felling must follow seeding".<br />

Local communities should be consulted about the timing <strong>of</strong> logging operations, and logging cycles<br />

should attempt to compliment the slack time in the agricultural cycle, so as to provide jobs for local<br />

people (Dykstra and Heinrich, 1992). Local people should also be infonned <strong>of</strong> logging operations and<br />

consulted on the timing with regards to the harvest <strong>of</strong> NTFPs. They should be allowed access to<br />

complimentarily harvest NTFPs prior to logging, as with the cutting <strong>of</strong> rattan 43 , the collection <strong>of</strong><br />

oil producing seeds and tapping <strong>of</strong> essential oils from valuable timber species such as Carapa<br />

guianensis and Copaifera species in the Amazon, or the collection <strong>of</strong> medicinal barks, such as those<br />

on the important timber Tabebuia species used to treat cancer, or Myroxylon balsamum in Peru.<br />

Following logging, local communities should be provided access to collect fuelwood and other<br />

products 44 •<br />

Harvest planning requires an initial increased expenditure, but cuts down on problems and costs by<br />

reducing wastage and increasing efficiency. Well-planned harvesting operations are not only<br />

preferable from a sustained use perspective, but have proven to be a great deal cheaper, as we11 45 •<br />

Road Construction<br />

In poorly planned logging operations, roads tend to occupy an inordinate portion <strong>of</strong> the forest. A.D.<br />

John (1992) found that roads and loading or landing areas occupied 6-20% <strong>of</strong> the forest area in<br />

Malaysia, and Uhl (1989) found roads to cover 8% <strong>of</strong> the logged forest in the eastern Amazon.<br />

Additionally, roads in tropical forest logging operations tend to lack effective drainage and sutface<br />

materials, so are susceptible to rain and traffic, resulting in > 90% <strong>of</strong> the soil erosion resulting from<br />

timber harvesting (Jonson and Lindgren, 1990; FAO, 1977). Dykstra and Heinrich (1992) find roads<br />

to be "unquestionably the most problematic feature <strong>of</strong> timber-harvesting operations". But competent<br />

technical supervision and planning can drastically minimize erosion and increase the retention <strong>of</strong><br />

forest cover. In North Queensland, the width <strong>of</strong> major extraction roads was limited to no wider than<br />

7.5 m, and for minor extraction roads 5.0 m; road grades were generally restricted to 8 degrees (14%)<br />

and 46-58 degrees (dependent on soil types) for sidecut roads (North Queensland Guidelines). This<br />

42 In North Queensland, no ttfelling <strong>of</strong> trees or snigging or hauling <strong>of</strong> logstl was allowed between 1 January and 31<br />

March in any year, except when drainage works were completed by 31 December in the previous year and weather<br />

conditions were suitable.<br />

43 In Sabah, Abdillah and Phillips found that logging resulted in a 76% reduction in the number <strong>of</strong> valuable<br />

commercial rattan and increased the number and size <strong>of</strong> patches devoid <strong>of</strong> rattan. The felling and extraction <strong>of</strong> timber<br />

accounted for 95% <strong>of</strong> this damage.<br />

44 In The Philippines, shifting cultivators settled recently logged-over forest and, utilizing income from the extraction<br />

<strong>of</strong> remaining timber and charcoal production, and applying knowledge <strong>of</strong> forest dynamics from their places <strong>of</strong> origin,<br />

developed a sustainable perennial tree-crop and root based agr<strong>of</strong>orestry system, reflecting an interactive process <strong>of</strong> change<br />

and adaptation between natural and human systems (Fujisaka and Wollenberg, 1991).<br />

45 The UNDP/FAO Forestry Development Project Sarawak found that improved forest management and harvesting<br />

resulted in a decrease in residual damage and a 20% reduction in costs. The Celos Management System in Suriname<br />

reported less residual damage to trees, less total disturbance <strong>of</strong> soil, and a 20-45% reduction <strong>of</strong> costs overall as a result<br />

<strong>of</strong> well-planned harvesting operations (de Graaf and Poels, 1990; Hendrison, 1990).


is in marked contrast to an average width <strong>of</strong> logging roads in Papua New Guinea reported by the<br />

FAO in 1989 as 18.4 m, and the average total deforested width <strong>of</strong> primary roads <strong>of</strong> 60 m in a logging<br />

concession in the Republic <strong>of</strong> Congo (Wilkie et al., 1992).<br />

Perhaps the largest impact <strong>of</strong> logging roads is the access they provide for small-scale loggers, hunters<br />

and fanner to once inaccessible land, timber trees and populations <strong>of</strong> wildlife (Davies, 1987;<br />

Caldecott, 1989; Dahaban, Nordin and Bennett, 1992). Guidelines for sustainable logging usually<br />

recommend the prompt closure or replanting <strong>of</strong> roads upon completion <strong>of</strong> the felling cycle. Many will<br />

deteriorate very quickly <strong>of</strong> their own accord due to the climatic extremes common in these regions 46<br />

(Redford, 1990).<br />

Felling Operations<br />

Selective felling <strong>of</strong> trees can mimic natural tree fall, but in reality is usually far more destructive. The<br />

extent <strong>of</strong> felling damage is largely determined by the felling intensity, or the number or volume <strong>of</strong><br />

trees felled per hectare. The Celos Management System found that a maximum <strong>of</strong> 5-8 trees per ha<br />

can be sustainably extracted from most tropical rainforests (Hendrison, 1990). Felling damage can<br />

also depend on the training and supervision <strong>of</strong> crews (for example, many crews do not know to make<br />

an undercut to direct the fall <strong>of</strong> a tree). With proper planing and measurement, the Celos Harvesting<br />

System crews felled 80% <strong>of</strong> trees in the desired direction, that is towards skid trails and gaps. Only<br />

11 % fell in an unfavourable position (Hendrison, 1990). Once down, proper cross-cutting <strong>of</strong> trees can<br />

maximize the value <strong>of</strong> the tree and reduce wastage. Manual pit-sawing, hand-held power chainsaws,<br />

or portable sawmills can produce boards and blocks, although these might be <strong>of</strong> lower quality than<br />

those produced in a mill; additionally, these systems do not require road construction and heavy<br />

machinery, so minimize damage to the forest (Dykstra and Heinrich, 1992; Jonson and Lindgren,<br />

1990; Hartshorn, 1990; Davies and Richards, 1991). The percentage <strong>of</strong> felled timber left unutilized<br />

in the forest is generally twice that in temperate regions. Combined with wastage during<br />

manufacturing, total timber utilization rates can be as low as 30% (Jonson and Lindgren, 1990).<br />

Cutting climbers prior to felling can also reduce damage and wastage by reducing the number <strong>of</strong> trees<br />

(many <strong>of</strong> the pole-sized classes that could form the next crop <strong>of</strong> commercial species) knocked over<br />

or broken by 25-50% (Dykstra and Heinrich, 1992; Jonson and Lindgren, 1990). Proper felling<br />

techniques can reduce the area logged each year to provide the current volume <strong>of</strong> tropical industrial<br />

roundwood.<br />

Skidding and Yarding Operations<br />

Most logging in tropical forests is characterized by heavy hauling and transporting machinery that<br />

is capital-rather than labour-intensive. This equipment, largely developed for other purposes, has<br />

become common to tropical logging operations since the 1950s, and has contributed to a 13-fold<br />

increase in the use <strong>of</strong> tropical hardwoods by the main importing countries (John, A.D.; Whitmore,<br />

1990). Conventional ground skidding equipment and extraction practices result in two major forms<br />

<strong>of</strong> damage: excessive damage to residual trees and advanced regeneration because skidders tend to<br />

wander through the forest in search <strong>of</strong> felled trees, which leads to a proliferation <strong>of</strong> skid trails; and<br />

46 In the Amazon, roads have <strong>of</strong>fered alternative transport for forest communities historically dependent upon the elites<br />

who violently controlled river transport, and so the trade in every forest commodity. If land title is assured for forest<br />

residents prior to the opening <strong>of</strong> roads, the movement <strong>of</strong> settlers and resulting unrestricted harvests <strong>of</strong> forest products can<br />

be somewhat minimized (Clay and Clement, 1993).<br />

23


24<br />

soil disturbance and compaction, which increases the potential for erosion and retards the growth <strong>of</strong><br />

regeneration (Dykstra and Heinrich, 1992)47.<br />

There also exist a number <strong>of</strong> potentially less-damaging alternative extraction systems. Cable systems<br />

reduce road needs and soil disturbance, but require highly skilled crews and are complicated and<br />

expensive to run. The ecological impact <strong>of</strong> manual extraction is minimal, but severely strains the<br />

laborer, and protective gear is rarely available. Animals have become a somewhat popular alternative<br />

to conventional mechanical skidding equipment. In Asia, elephants have long been used to extract<br />

logs, and in Latin America and Asia, bullocks and oxen are employed. The advantages <strong>of</strong> this system<br />

are the minimal ecological disturbance, high flexibility and adaptiveness, low investment costs, and<br />

intensive labour requirements, which results in the employment <strong>of</strong> more local people. The<br />

disadvantages include the need to care for the animals even when they are not being used, the loss<br />

<strong>of</strong> efficiency when harvesting logs > 70 cm dbh, and the relatively slow pace compared with<br />

mechanical equipment. In Palcazu, Peru, it was found that the positive attributes <strong>of</strong> this system far<br />

outweighed the negative (Ocana-Vidal, 1992; Hartshorn, 1992). Other alternative forms <strong>of</strong> extraction<br />

include railways, helicopters, and balloons, although none have caught on to any great extent (Jonson<br />

and Lindgren, 1990).<br />

Unlike tree fall gaps, mechanized transport <strong>of</strong> felled trees over the forest soil has no natural<br />

equivalent and damages the flora and soil beyond anything comparable to damage caused by large<br />

herbivores inhabiting the forest or by fallen trees sliding downhill (Hendrison, 1990). As a result, the<br />

objective <strong>of</strong> harvesting operations can only be to minimize negative impacts on both the potential<br />

timber and the many existing and potential non-timber values provided by the forest in which it is<br />

carried out.<br />

Post-harvest Assessments<br />

Post-harvest assessments determine the level to which a harvesting plan has been implemented and<br />

management objectives achieved. They are critical for sustainable forest management, as harvest<br />

plans are not <strong>of</strong>ten followed in spirit nor in practice. These assessments include: a report on operation<br />

costs and revenues, an evaluation <strong>of</strong> the degree to which silvicultural, non-timber product, and<br />

biodiversity conservation objectives are met, residual stand characteristics and the extent <strong>of</strong> damage,<br />

area disturbed by roads and skid trails, the quality and quantity <strong>of</strong> regeneration, and the extent <strong>of</strong><br />

revegetation and closing <strong>of</strong> skid trails, landing sites and logging roads. The results <strong>of</strong> these<br />

assessments must be shared with crews, and financial incentives or penalties imposed for the quality<br />

<strong>of</strong> the work (Dykstra and Heinrich, 1992; de Graaf and Poels, 1986; Jonson and Lindgren, 1990).<br />

47 But damage caused by ground skidders in combination with bulldozers, by far the most common method <strong>of</strong><br />

extraction, can be minimized. Low ground pressure tracked skidders cause less damage to the soil than wheeled skidders.<br />

Pre-planned skid trails, directional felling toward skid trails, and winching <strong>of</strong> logs from the stump 10 the skid trail when<br />

possible, can lead to a reduction in residual forest damage, in the number <strong>of</strong> merchantable logs left in the forest, and can<br />

reduce costs by a third (Dykstra and Heinrich, 1992; Jonson and Lindgren, 1990; Hendrison, 1990). The Celos Harvesting<br />

System, incorporating these practices, resulted in twice the production <strong>of</strong> conventional skidding, 8% damage to residual<br />

forest in the extraction <strong>of</strong> 8-10 trees (as compared with 14% resulting from conventional felling and skidding operations),<br />

and 40% less time required per unit product transported (Hendrison, 1990).


Box 3: Inventories <strong>of</strong> Non-Timber Forest Products for Natural Forest Management<br />

An inventory <strong>of</strong> NTFPs should provide a reasonably precise estimate <strong>of</strong> the total number <strong>of</strong> halVestable<br />

trees per hectare. For fruit and oil species, this means the total number <strong>of</strong> adult trees, for latexproducing<br />

species, medicinal plants and rattan, some juvenile trees may also be included. Before<br />

initiating field work, the merchantability limits <strong>of</strong> the resource should be established (peters, 1993).<br />

Assessment <strong>of</strong> NTFPs conducted in Ghana varied by plant type. For climbers, the number <strong>of</strong> stems is<br />

counted, and for herbaceous species the number <strong>of</strong> plant groups (clumps). For cane, the numbers <strong>of</strong><br />

mature, immature, and cut stems are recorded; this will provide a crude indication <strong>of</strong> plant densities and<br />

exploitation levels (Falconer, ODA, 1992).<br />

The inventory should also provide data on the current population structure or size-class distribution <strong>of</strong><br />

the species. The minimum diameter limit will be far below that for timber species, and will vary<br />

depending upon the size and abundance <strong>of</strong> species. For species that occur at relatively low density, 10<br />

cm dbh may be reasonable; more abundant species will probably require a slightly higher diameter cut<strong>of</strong>f<br />

(peters, 1993; Falconer, 1992; Boom, 1989). Tree sampling methods should be modified to<br />

incorporate species <strong>of</strong> economic importance that are halVested when small. In Ghana, for example,<br />

Garcinia epunctata is harvested at between 5-10 cm dbh for chewsticks (Falconer, 1992). In conducting<br />

an inventory <strong>of</strong> xate, allspice and chicle in Guatemala, Reining et al. (1992) set minimum diameter<br />

limits at 10 cm dbh for all species except allspice, which was measured down to 5 cm dbh. Gentry<br />

(1992) suggests that in the study he conducted with Peters (1989), in which they attempted to document<br />

the actual value <strong>of</strong> a single hectare <strong>of</strong> Amazonian forest, their values were probably underestimated<br />

because they did not include species utilized for fibre and medicine that are usually harvested from<br />

plants < 10 cm dbh (Gentry, 1992).<br />

A predetermined percentage <strong>of</strong> the area is sampled using plots or transects (<strong>of</strong> square or circular<br />

configuration), stratified among different forest types found within the area. NTFP inventories must<br />

<strong>of</strong>ten be incorporated into on-going inventory work. In Ghana, the NTFP inventory was forced for<br />

logistical reasons to use an existing sampling design <strong>of</strong> 1 ha sample plots stratified systematically.<br />

Temporary sample plots were established in forest logged within the last three years and in unlogged<br />

forest to assess the effects <strong>of</strong> varying intensities <strong>of</strong> logging on NTFPs (Falconer, 1992).<br />

Permanent sample plots, which can provide data on changes over time, should be established to monitor<br />

the impacts <strong>of</strong> non-timber and timber extraction on forest products. Ideally, plots should be random and<br />

stratified by forest type (pairs <strong>of</strong> 1 ha each), and probably <strong>of</strong> square or broadly rectangular shape to<br />

minimize edge effects and because they are faster and easier to demarcate than circular plots in most<br />

tropical forests. Plots will be periodically re-inventoried, usually every five years (Alder and Synnott,<br />

1992; Reining et al., 1992).<br />

The exact number <strong>of</strong> PSPs used will depend upon the abundance <strong>of</strong> populations - high-density<br />

populations will require fewer plots than scattered, low-density populations, which must be more<br />

intensively sampled (peters, 1993). The ODA Forest Resources Management Project in Ghana has<br />

established 629 100 m x 100 m PSPs to provide information on the NTFP growing stock and to<br />

measure the growth <strong>of</strong> individual trees in different forest types. In two sub-plots <strong>of</strong> each, trees will be<br />

measured down to 1 cm dbh so as to collect data on the species harvested at small sizes.<br />

25


26<br />

In this way, regeneration <strong>of</strong> species commercially harvested at much larger diameters can also be<br />

assessed, and the intensity <strong>of</strong> halVesting practices reduced if the density <strong>of</strong> seedlings and saplings are<br />

found to drop. The effectiveness <strong>of</strong> this harvest reduction will be verified and adjusted, if necessary,<br />

during the next inventory. (peters, 1993; Falconer, 1992; Salick, 1992). In the Si-a-Paz International<br />

Peace Park in Nicaragua and Costa Rica, stratified random subplots, totalling 10% <strong>of</strong> the PSPs<br />

established to measure regeneration, are used to inventory and voucher, with the help <strong>of</strong> a local,<br />

knowledgeable infonnant, useful plants and their community characteristics (species richness, diversity,<br />

density and cover) (Salick, 1992). The interview/inventory technique, involving the collection <strong>of</strong> plant<br />

specimens and the subsequent interviewing <strong>of</strong> informants as to names and uses, was employed by Boom<br />

(1989) in his study <strong>of</strong> the ethnobotany <strong>of</strong> the Chacabo Indians in Bolivia. PSPs can also be the subject<br />

<strong>of</strong> detailed botanical inventories, which record all plants found on the plot (Falconer, 1992; Salick,<br />

1992).<br />

VII. Silvicultural Systems for Natural Forest Management<br />

Silviculture and harvesting systems presuppose social, economic and ecological knowledge, as defined<br />

in management objectives (Schmidt, 1991). The characteristic tools <strong>of</strong> silviculture include the<br />

regulation <strong>of</strong> shade and canopy opening, treatments to promote valuable individuals and species and<br />

reduce unwanted trees, climber cutting, "refining", poisoning, enrichment, and selection (Poore et al.,<br />

1989).<br />

All silvicultural systems will, to some extent, change and usually coarsen the fine mosaic <strong>of</strong> gap,<br />

building and mature phases <strong>of</strong> a natural forest (Whitmore, 1991). Silvicultural systems that aim to<br />

maximize the variety <strong>of</strong> products extracted from the forest and retain species diversity should change<br />

the natural composition and structure <strong>of</strong> the forest as little as possible (Poore and Sayer, 1991;<br />

Parren, 1992). As the ITIO Guidelines for Sustainable Forest Management (1990) suggest: "the<br />

choice <strong>of</strong> silvicultural system should be aimed at sustained yield at minimum cost, enabling<br />

harvesting now and in the future while respecting recognized secondary objectives". In many cases,<br />

non-timber objectives will not be considered secondary and will have an economic and social value<br />

<strong>of</strong> equal or greater importance than timber; silvicultural systems should be selected and adapted to<br />

reflect this value.<br />

Silvicultural systems for natural forests are grouped into the polycyclic and monocyclic. Monocyclic<br />

systems remove all saleable trees from a forest at a single operation, resulting in bigger gaps in the<br />

canopy best suited for the growth <strong>of</strong> pioneer species with light, pale marketable timber (Whitmore,<br />

1991). Monocyclic systems dramatically alter forest structure and species diversity, and so severely<br />

limit any long-tenn harvesting <strong>of</strong> non-timber forest products.<br />

Polycyclic systems are usually more conducive to conservation and non-timber objectives, although<br />

it must be said that neither system truly reflects multiple-use management objectives that specifically<br />

promote the sustainable harvest <strong>of</strong> non-timber, as well as timber, products (Nair, 1991). Polycyclic<br />

systems result in scattered gaps in the canopy, favouring slower growing shade-bearers, and relying<br />

on the release <strong>of</strong> advanced regeneration <strong>of</strong> commercial adolescents on the forest floor. Under the best<br />

conditions, these systems can increase the timber yield over a full rotation, but this depends largely<br />

on the intensity <strong>of</strong> and residual damage incurred by timber harvesting, subsequent competition from


colonizers, and the success <strong>of</strong> advance regeneration that survives the microclimate changes resulting<br />

from canopy opening (Nair, 1991; Whitmore, 1992). "High grading" or "creaming" <strong>of</strong> commercial<br />

species that display desirable traits can result in disgenic effects in polycyclic systems, and must be<br />

carefully controlled (Hendrison, 1990; Queensland Guidelines).<br />

Natural forest management systems can be extensive or intensive. Intensive management has proved<br />

extremely problematic in the multi-age class and multi-species tropical moist forests due to our<br />

ignorance <strong>of</strong> the ecology <strong>of</strong> commercial species and the ecosystems in which they operate (Poore,<br />

1989). Extensive systems range, with increasing productivity and costs, from the "wait and see"<br />

demarcation <strong>of</strong> forests, to the "log and leave", to the minimum intervention, stand treatment and<br />

enrichment planting <strong>of</strong> low intensity natural regeneration silvicultural systems (Poore, 1989; Jonson<br />

and Lindgren, 1990). With increasing intensity in forest management, there exists a trade-<strong>of</strong>f with<br />

the conservation and non-timber values associated with a forest. "Wait and see" and "log and leave"<br />

forests (providing logging damage is minimized) will provide the most opportunity for the<br />

conservation <strong>of</strong> forest structural and species diversity and the variety <strong>of</strong> products resulting from this<br />

diversity. But in many forests timber production will be a primary management objective. Low<br />

intensity management, utilizing polycyclic, natural regeneration silvicultural systems, produces larger<br />

volumes <strong>of</strong> timber, while - to varying degrees - maintaining non-timber values. Low intensity,<br />

polycyclic management <strong>of</strong> natural forests can be broken down into three, more or less distinct,<br />

systems: light selective logging, light selective logging and silvicultural treatment, and light selective<br />

logging and enrichment planting 48 •<br />

Light Selective Logging or "Minimum Intervention"<br />

This is the simplest system, relying on silvicultural knowledge and using models to define the<br />

sustainable harvesting intensity, girth limits, and cutting cycles. Only stems <strong>of</strong> marketable species are<br />

removed, no other species being interfered with. Natural regeneration fills the gaps created by felling,<br />

without manipulation or enrichment. Forests managed in this way must be sufficiently stocked with<br />

timber trees <strong>of</strong> large dimensions to make utilization pr<strong>of</strong>itable. Management includes inventories and<br />

harvest planning. Upon completion <strong>of</strong> felling operations, the forest is closed until the next cycle.<br />

Although the growth rates <strong>of</strong> commercial species "managed" in this way are <strong>of</strong>ten considered<br />

uneconomically low, light selective logging, or "light creaming", is still widely practised in the<br />

tropics, can be integrated into existing NTFP harvesting practices, and is one <strong>of</strong> the least destructive<br />

land use systems practised in the tropics, provided responsible harvesting practices are employed<br />

(Gomez-Pompa and Burley, 1991; Poore, 1989; Jonson and Lindgren, 1990; Hendrison, 1990).<br />

Light Selective Logging and Stand Treatment<br />

In these systems logging is carried out as described above, but is followed by treatments to "direct"<br />

natural regeneration to increase the representation <strong>of</strong> commercial species. This includes the poisongirdling<br />

and weeding <strong>of</strong> undesirable species, and the cutting <strong>of</strong> competing lianas and brush, but does<br />

not include direct measures to change the forest structure (Jonson and Lindgren, 1990). Because<br />

tropical moist forests contain a mix <strong>of</strong> tree species, "refining" is employed to eliminate the<br />

undesirable competitors <strong>of</strong> commercial species. Competition is reduced by: liquidating all non-<br />

48 These types <strong>of</strong> silvicultural systems attempt to answer the following questions: are there sufficient seedlings,<br />

saplings, and advanced growth <strong>of</strong> merchantable species at the time <strong>of</strong> exploitation to provide adequate stocking for the<br />

next crop? what are the silvics <strong>of</strong> these species (most importantly their regeneration potential) and what ·treatment will<br />

be necessary? and what are the probable rates <strong>of</strong> growth and merchantable volume expectations <strong>of</strong> the different species?<br />

(Wyatt-Smith, 1987).<br />

27


28<br />

desirable species (and defective individuals <strong>of</strong> desirable species) above a planned girth limit;<br />

individual release <strong>of</strong> "leading desirables"; and liquidating non-desirables starting with the biggest<br />

trees downwards until a predetermined residual basal area is attained (van der Hout, 1992)49.<br />

It is difficult to detennine the change in forest productivity due to refinement, since growth<br />

variability is so extreme, and is influenced by so many interacting factors that it is "sufficiently<br />

random to frustrate any attempt to interpret it in detenninistic tenns" (Wadsworth, 1987; Schmidt,<br />

1991). "Sustainability" cannot be detennined before the third felling cycle, and this type <strong>of</strong> data is<br />

not currently available (Nair, 1991). Refining is also a direct attack on the ecological diversity <strong>of</strong><br />

forests, and may gradually eliminate more than half the tree species. De graaf and Poels (1990) admit<br />

ignorance <strong>of</strong> the ecological effects <strong>of</strong> reductions in the numbers <strong>of</strong> non-commercial species.<br />

Wadsworth (1987) suggests that developing additional uses for the existing forest stock would be<br />

preferable to attempts to induce something that does not already exist in the forest - that is, try to<br />

do through processing and marketing what is attempted through poisoning (although the harvest <strong>of</strong><br />

larger numbers <strong>of</strong> species creates the danger <strong>of</strong> increased logging intensity).<br />

The less tampering done with the undergrowth, top soil and microclimate <strong>of</strong> the forests, the easier<br />

it is to maintain the ecosystem commercial trees are dependent upon (Dawkins, 1958; Meijer, 1970).<br />

Meijer (1970) found that girdling the undergrowth in dipterocarp forests could enhance conditions<br />

for the invasive nomad trees for 20-40 years. It has also been found that species deemed<br />

noncommercial, and so "refined", have later proved to be important keystone species, or have<br />

exhibited significant economic value, if not a greater value than those their absence has liberated<br />

(Schmidt, 1991; Hammond and Brown, 1991; ITTO, 1992). In the Celos Management System,<br />

commercial species are "defined on market criteria and the technological qualities <strong>of</strong> the wood". In<br />

the forest area in Suriname where they work, this amounts to only 40-50 species (de Graaf and Poels,<br />

1990; 1991). There is obviously little room in this system for the development <strong>of</strong> markets for<br />

additional timber species or any long-term non-timber forest product extraction.<br />

But the CMS, and other systems like it, are attempting to create a fine balance between what is<br />

considered to be the uneconomically slow growth <strong>of</strong> commercial species in natural forests without<br />

silvicultural treatment, and the need to retain forest structural and species diversity. Many find that<br />

for timber production alone, logging without silvicultural treatments is not sufficiently economical<br />

(de Graaf and Pools, 1990; Silva et al., 1992; Chelunor Nwoboshi, 1987). In many areas, however,<br />

the economic and social importance <strong>of</strong> timber is equal or even secondary to that <strong>of</strong> non-timber<br />

products, and silvicultural treatments that coarsen forest structure and greatly impact forest diversity<br />

should be abandoned. Whitmore (1992) notes that in large part post-harvesting treatments have been<br />

abandoned due to their expense, and that growth rates can be improved through the control <strong>of</strong> logging<br />

operations. Thus, timber felling, usually regarded as the first major silvicultural operation in a natural<br />

regeneration system, has become the only canopy manipulation the forester can afford (Wyatt-Smith,<br />

1987; Whitmore, 1992).<br />

49 Refinements are expected to reduce the cutting cycle from 60 to 30 years in Sarawak, and the shifts in species<br />

composition it produces have been credited with 50-250% gains in yield (van der Hout, 1992; Wadsworth, 1987). De<br />

Graaf and Poels (1986; 1990) found that the reduction <strong>of</strong> non-commercial species by refinement (which eliminated 1(1­<br />

2/3 <strong>of</strong> the standing stem volumes <strong>of</strong> these species) produced an increase in annual production <strong>of</strong> commercial timber to<br />

2m 3 /ha from 0.5 m 3 /ha. Wadsworth (1987) found that refinements most significantly created gains in quality, and not in<br />

increased wood volume.


Climber cutting is a perhaps less controversial issue in natural forest management. Climbers not only<br />

smother regeneration in logged over forest but, by binding trees together, can result in the destruction<br />

<strong>of</strong> large groups <strong>of</strong> trees during felling operations. It has been found that climber cutting, at least a<br />

year in advance <strong>of</strong> logging operations, reduced the number <strong>of</strong> trees pulled down during felling by 1/4<br />

(Jonson and Lindgren, 1990)50.<br />

Many climbers play an important role in local and international economies, and these should be<br />

harvested prior to the felling <strong>of</strong> timber, and any endangered species carefully avoided. Rare<br />

Ancistrocladus climbers in Cameroon, for example, have yielded potential anti-HIV compounds,<br />

which could generate significant economic returns should a pharmaceutical be developed (Katz­<br />

Miller, 1993). The Amazonian liana Chondodendron tomentosum, used in curares, or arrow poisons,<br />

<strong>of</strong> Amerindians in Colombia, Ecuador and Peru, was found to contain the valuable skeletal muscle<br />

relaxant tubocurarine (Schultes, 1992). Amazonian shaman <strong>of</strong>ten cultivate the wild liana<br />

Banisteriopsis caapi to prepare the hallucinatory ayahuasca for ceremonies (Schultes, 1992). Gnetum<br />

bulchozi and Heinsia crinita are sources <strong>of</strong> important spinach-like greens in diets in West Africa<br />

(Agwu, pers. comm., 1993). Climbing palms are also important in this region for household and<br />

commercial weaving, including Eremospatha spp., Laccosperma opacum and Calamus deeratus<br />

(Falconer, 1992).<br />

The neotropical Desmoncus species have shown similar growth habits to the valuable Calamus<br />

species <strong>of</strong> rattan, are used in basket-making by indigenous peoples, and have been suggested as<br />

potential substitutes to Southeast Asian rattan (Gentry, 1992; Clay and Clement, 1993)51. Recently<br />

a cottage industry in rattan products has sprung up around Iquitos, Peru, where there is already a<br />

thriving fibre industry based on the aerial roots <strong>of</strong> Philodendron solimiesensis, which have long been<br />

extensively used in campesino handicrafts (Gentry, 1992). Many <strong>of</strong> these fibre plants occur in<br />

second-growth forest, and their utilization could generate substantial incomes locally. Fevillea spp.<br />

have seeds richer in oil than any other dicot, and Amerindians in Peru use them as candles (Gentry,<br />

1992).<br />

Light Selective Logging and Enrichment Planting<br />

In these systems, saplings <strong>of</strong> desired species are planted in lines or patches where stocking <strong>of</strong> residual<br />

stems is low. Species planted are usually rapid-growing, and can result in the gradual elimination <strong>of</strong><br />

existing stands (Wadsworth, 1983). In Venezuela, researchers have found a more than 30% drop in<br />

species in strips planted with non-indigenous timber. The impact <strong>of</strong> enrichment planting was<br />

particularly pronounced on specialist species (Grajal, pers. comm., 1993). The IITO Draft Guidelines<br />

for Biological Diversity Conservation in Forests Managed for Timber (1992) and The Global<br />

50 Following logging, the greater sprouting capacity <strong>of</strong>climbers, their competitive investment in resource procurement<br />

surfaces (relying on the trees' supportive tissues), and the climbing structures abundant in the slash <strong>of</strong> logged-over forest,<br />

<strong>of</strong>ten lead to the smothering <strong>of</strong> seedlings, damage to poles, and the formation <strong>of</strong> climber tangles through which<br />

regeneration has difficulty emerging (Fox, 1968; UbI et al., 1981). Pioneers might be better suited to survive climber<br />

infestations than climax species, as they are <strong>of</strong>ten flexible and fast growing, and have large compound leaves or leaf-like<br />

branches which may help them shed lianas (putz, 1984).<br />

51 The high economic value <strong>of</strong> rattan in Southeast Asia have resulted in organized teams (not always made up <strong>of</strong> local<br />

people, who may not be informed <strong>of</strong> the timing <strong>of</strong> logging operations or as aware <strong>of</strong> their implications as outside<br />

merchants and traders) extracting rattan prior to logging (Abdillah and Phillips; Stockdale, pers. comm., 1993).<br />

29


30<br />

Biodiversity Strategy (1992) recommend that enrichment planting use native wild seedlings or<br />

seedlings raised from locally collected sources. Recent research on enrichment planting <strong>of</strong> native<br />

species has suggested that, contrary to existing beliefs, native species can <strong>of</strong>ten be more productive<br />

than the exotics which replace them although in some cases these plantings could suffer from<br />

predation due to changes in density-dependent plant-herbivore relations «WRI et al., 1992; Hartshorn,<br />

1980).<br />

Enrichment plantings <strong>of</strong> multi-purpose species, such as Carapa guianensis (an important timber<br />

species in the Amazon, the seeds <strong>of</strong> which yield a valuable medicinal oil), Caryocar villosum (an<br />

important timber and fruit tree in the Amazon), and Brazil nut (Bertholletia excelsa) has been tried<br />

in some areas <strong>of</strong> the Amazon (Clay and Clement, 1993). Traditional agr<strong>of</strong>orestry and forest<br />

management systems <strong>of</strong>ten employ enrichment plantings <strong>of</strong> desirable species (Alcorn, 1990; Gomez­<br />

Pompa, 1990; Balee, 1989). Clement (1993) suggests that enrichment planting with NTFP species<br />

could have the added benefit <strong>of</strong> introducing high-quality gennplasm into the forest. Combined with<br />

natural regeneration strategies, this could lead to an increase in yield and market quality. Gennplasm<br />

collections are expensive to develop, maintain, characterize and evaluate correctly, so traditional<br />

home gardens would probably provide the best varieties for enrichment (Clay and Clement, 1993).<br />

Clearing Systems and NTFPs<br />

In addition to the low intensity systems described above, some <strong>of</strong> the more intensive clearing systems<br />

(as defined by Gomez-Pompa and Burley, 1991), promote the natural regeneration <strong>of</strong> valuable<br />

commercial species and eliminate the undesirables; future forests are thus enriched with propagules<br />

coming only from the commercial species. Mature trees, seedlings and juveniles <strong>of</strong> most undesirable<br />

species are eliminated. As a result, the forest is converted from a multi-species, multi-aged to a more<br />

or less even aged forest with a greater percentage <strong>of</strong> commercial species.<br />

Although they utilize natural regeneration, these systems have a strong affinity to conversion and<br />

leave little room for the harvest <strong>of</strong> a significant variety <strong>of</strong> non-timber forest products (van der Hout,<br />

1992; Chiew Thang, 1987). Wyatt-Smith (1987) remarked that an example <strong>of</strong> this system, the<br />

Malaysian Unifonn System, produced areas that resemble Shorea plantations. These systems have<br />

proven extremely difficult to implement because: the soil seed bank produces numerous undesirable<br />

species, which require expensive cleaning and weeding operations; the commercial species may not<br />

flower for several years; the opening <strong>of</strong> the overstorey to favour rapid growth also stimulates weeds<br />

and climbers; and knowledge <strong>of</strong> the growing stock, its distribution by species, size classes and<br />

location and how these change with harvesting and treatment is extremely limited. (Gomez-Pompa<br />

and Burley, 1991; Wadsworth, 1987; Schmidt, 1991). After 38 years <strong>of</strong> effort to regenerate forest in<br />

Nigeria under the Tropical Shelterwood System, failure was admitted and the project terminated<br />

(Wadsworth, 1987).<br />

However, the strip clear-cutting, or strip shelterbelt, system employed by the Yanesha Indians in the<br />

Palacazu Valley in eastern Peru, has proved promising for natural forest management, and efforts<br />

have been undertaken to incorporate non-timber forest products into forest management (Salick,<br />

1992). The project is attempting to find markets for a variety <strong>of</strong> wood products and a large number<br />

<strong>of</strong> species, while promoting vertical integration <strong>of</strong> processing industries. Long (200-500 m), narrow<br />

(30-40 m wide) strips are clear cut, and rotated through the forest in 30-40 year felling cycles in such<br />

a way that the undisturbed forest provides a source <strong>of</strong> propagules for regeneration (Hartshorn, 1990;<br />

Gomez-Pompa and Burley, 1991; Jonson and Lindgren, 1990). Buffer zones are retained along


watercourses, steep slopes, and smaller reserved areas are set aside every six strips. Disturbance by<br />

weeds, climbers and pioneers, and the removal <strong>of</strong> nutrients stored in forest vegetation could influence<br />

the rotation and species composition <strong>of</strong> the forest, but natural regeneration <strong>of</strong> commercial species<br />

from seed and coppice appears to be "sticking" (Hartshorn, 1990; Jonson and Lindgren, 1990).<br />

Gentry (1992), found that the current approach to harvesting does not yet make the most <strong>of</strong> the<br />

forest's diversity. Small trees (and branches) are used for charcoal, medium-sized trees for posts and<br />

telephone poles, and large trees for timber. Thus, in essence, "the diverse forest is treated as though<br />

it consisted <strong>of</strong> three species", and the potentially higher value <strong>of</strong> individual species as fruits,<br />

medicines, nuts, latexes, ete. is lost (Gentry, 1992).<br />

Salick (1992) has undertaken a study to determine the number <strong>of</strong> plants regenerating in a cleared<br />

experimental strip that are useful to the Yanesha. Overall, the number <strong>of</strong> useful plants and plant<br />

species increased dramatically, due to the small size <strong>of</strong> individuals and changes in species<br />

composition resulting from the early successional nature <strong>of</strong> the strip. In some cases, however, some<br />

<strong>of</strong> the herbs and small grasses collected were found to be rare, and may have unusual gap niches<br />

(Salick, 1992). Some species <strong>of</strong> importance to Yanesha subsistence were lost, however, and there is<br />

no evidence that they will regenerate in the near future. These include medicinal, construction<br />

materials, arrow poison, thatch and - most importantly - Heteropsis and Maregravia spp. used to<br />

weave baskets, mats, and in the construction <strong>of</strong> houses, fish traps, and baby hammocks. These species<br />

are becoming increasingly scarce in this area, and must be collected as much as a day's walk away<br />

(Salick, 1992)52. Salick (1992) recommends that Yanesha indigenous management practices and their<br />

use <strong>of</strong> secondary forest products be integrated into strip management.<br />

Polycyclic, selection systems relying on natural regeneration have many limitations. Nair (1991)<br />

reports that there exists no case in which such a system has been applied over a number <strong>of</strong> cycles<br />

continuously. Management plans are continually revised to include new areas for timber extraction<br />

and reduce felling cycles and girth limits. Because <strong>of</strong> improved accessibility, the increase <strong>of</strong> lesser<br />

known species with markets, and the acceptability <strong>of</strong> low girth logs for industrial use, Nair (1991)<br />

believes that selective logging is in danger <strong>of</strong> being replaced by clear-felling, more intensive systems.<br />

Dawkins (1958), while describing selection systems as the ideal form <strong>of</strong> management for forests<br />

which are little understood and where violent conversions are liable to cause irreversible and<br />

unfavourable changes in forest ecosystems, found selection systems difficult to manage and<br />

52 Salick (1992) found that the regenerating forest included fewer species <strong>of</strong> palms and fewer individual ferns than<br />

the original vegetation, whereas more species and individuals <strong>of</strong> herbs and grasses were found. The post-harvest treatment<br />

<strong>of</strong> cutting and levelling slash apparently encourages the increased dominance <strong>of</strong> vines throughout the strip. Strip<br />

regeneration is different from natural gap fonnation, and resembles swidden fallows in the proportion <strong>of</strong> secondary plant<br />

species encountered and the importance <strong>of</strong> coppicing as a source <strong>of</strong> regeneration.<br />

31


32<br />

uneconomical 53 • But improvements in ecological knowledge, logging practices and an increase in<br />

the use <strong>of</strong> lesser-known-species have been shown to alleviate many <strong>of</strong> the problems associated with<br />

these systems.<br />

Most importantly, well-run selection systems are the only systems that serve natural forest<br />

management objectives geared towards producing the best total return from all forest products<br />

without depleting the resource base (although the strip-clearing cutting system described could have<br />

promise in some areas)54. Not all forests designated for timber production should be under lowintensity<br />

management, but the current unpredictability <strong>of</strong> both forest ecosystems and markets for<br />

forest products argues for a significant portion <strong>of</strong> the forest reserves <strong>of</strong> a country to be flexibly<br />

managed for a wide range <strong>of</strong> timber and non-timber values (Poore and Sayer, 1991; ITTO, 1992;<br />

ITIO, 1990).<br />

53 As summarized by Wadsworth (1987), Dawkins' argument against selection systems is that: the removal <strong>of</strong> 10-12<br />

mature trees/ha may be economically marginal; the yield <strong>of</strong> 3.5 m 3 /ha/year is far less than that <strong>of</strong> plantations; the cut<br />

destroys 20-250/0 <strong>of</strong> the adolescent and pole stocks and damage tends to be progressive; longer cycles mean a larger<br />

harvest but with more damage and fewer trees remaining for future crops; shorter cycles reduce the yield/cut and increases<br />

the damage per area per unit <strong>of</strong> time; large crown diameter/dbh ratios (20+) required for rapid-growing species must<br />

develop early but cannot be obtained beneath larger trees; and yield prospects, where there is no market for intermediatesized<br />

trees, are no more than 1.4 m 3 /ha/year.<br />

54 As Dawkins (1958) said himself, "where forest commerce is primitive and forest science underdeveloped, the<br />

conservative course is to retain as many existing trees over as wide an area as possible, for fear that what is lost can never<br />

be replaced".


Box 4:<br />

Examples <strong>of</strong> Multi-purpose Amazonian Species with Present and Future Potential for<br />

Natural Forest Management<br />

Brazil nut (Bertholletia excelsa) is an ideal multipurpose species. It is light-demanding and could be<br />

planted in larger gaps following logging operations. It yields nuts after 15-20 years and timber after 50­<br />

100 years. Mature trees can yield 100-225 kg <strong>of</strong> unshelled nuts in a good year. Although it is illegal to<br />

fell a Brazil nut tree, large numbers continue to be cut for its fine timber (Smith et al., 1992). Brazil<br />

nut grows quickly with little vulnerability to pests and diseases, and with little intervention. Nuts are<br />

harvested in the rainy season, while rubber, its compliment in extractive systems, is tapped during the<br />

dry season (when timber harvesting operations also take place) (Clay and Oement, 1993; Balee, 1989;<br />

Mori, 1992; Richards, 1993).<br />

Andiroba (Carapa guianensis) is a fast growing moderate shade-bearer. Andiroba is widely distributed<br />

in the neotropics and is frequently found in the Amazon basin in association with Virola surinamensis<br />

and Hevea brasiliensis, both mUlti-purpose species (Sampaio in: Clay and Clement, 1993; Tropical<br />

Woods, vo!. 90, 1947). The wood <strong>of</strong> andiroba is considered one <strong>of</strong> the Amazon's fmest (Record and<br />

Mell, 1924). The seeds <strong>of</strong> andiroba produce an extremely bitter oil (its name derives from the Indian<br />

words "nhandi" (oil) and "rob" (bitter). Small quantities <strong>of</strong> oil are used to sooth muscular distentions,<br />

skin tumours and superficial skin ailments. The Indians <strong>of</strong> French Guiana extract oil <strong>of</strong> the andiroba<br />

seed, mix it with Bixi oreallana, and prepare an ointment applied to the skin to prevent mOSQuito and<br />

flea bites. Andiroba is an ideal species for enrichment planting following logging. Its seed oil could<br />

provide annual income after the tenth year until the trees are large enough to cut for timber at between<br />

18-23 years (Sampaio, 1993 in: Clay and Clement, 1993).<br />

Andiroba is endangered in its natural habitat, and was proposed for inclusion in Appendix 11 <strong>of</strong> the<br />

Convention on International trade in Endangered Species (CITES) (Wellner and Dickey, 1991).<br />

Copaiba (Copaifera multijuga) is a large tree that can attain 36 m in height with a dbh <strong>of</strong> up to 80 cm;<br />

average dbh is between 40-50 cm (Alencar, 1981). It generally occupies the upper canopy and may<br />

occasionally be an emergent. Like other upper canopy and emergent species, C. multijuga requires<br />

shade during the seedling stage but requires sun in order to attain height and girth (Sampaio in: Clay<br />

and Clement, 1993). C. multijuga is widely used for sawn lumber, construction and carpentry, and<br />

makes a good charcoa1 55 • Oil resin, collected by drilling a hole in the trunk <strong>of</strong> C. multijuga, C.<br />

reticulata, C. guianensis and C. <strong>of</strong>ficinalis has large regional and international markets. It is used as a<br />

component <strong>of</strong> high temperature resistant varnishes, as a perfume fixative in cosmetics, antibacterial in<br />

creams and soaps, as a substitute for linseed oil in paints, as a modifying agent in plastics, and to<br />

improve the clarity <strong>of</strong> the image in low-contrast areas in photographic film development. The oil resin<br />

is also a popular medicine in Amazonian. It is used to treat throat infections, bronchitis and other<br />

respiratory problems, as an antiseptic for wounds and scratches, and as a cure for diarrhoea and<br />

problems <strong>of</strong> the urinary tract. Controlled tapping <strong>of</strong> the oleo-resin <strong>of</strong> Copaijera species could yield a<br />

continual source <strong>of</strong> sustainable income throughout the growth cycle <strong>of</strong> the tree (Shanley, 1993; Sampaio<br />

in: Clay and Clement, 1993).<br />

55 The wood is heavy, with a regular grain and medium texture similar to that <strong>of</strong> Cedrella odorata (Sampeio in: Clay<br />

and Clement, 1993). Copaijera species can produce excellent export veneer with a very attractive gain. Although it is<br />

relatively uncommon, generally <strong>of</strong> small diameter and difficult to harvest, Copaijera spp. have been cut down at<br />

unsustainable rates in the Peruvian Amazon (Gentry and Vasquez, 1988).<br />

33


34<br />

Piquia (Caryocar villosum) is one <strong>of</strong> the largest forest trees in most <strong>of</strong> its distribution, frequently<br />

attaining 40-50 m when occurring as an emergent above the canopy. Piquia does not regenerate well in<br />

the shade <strong>of</strong> the high forest, but growth is rapid when released by increased light (Clement, 1993).<br />

Piquia represented 1.1% (26,540 m 3 ) <strong>of</strong> the timber commercialized in Manaus in 1972. The wood is<br />

used for ship-building, civil construction, and general carpentry. Amerindians rely heavily on its fruit,<br />

and caboclos know the location <strong>of</strong> most trees and visit them frequently during harvest season. Piquia<br />

could be planted in the larger gaps left after logging (Clay and Clement, 1993). Piquia is endangered;<br />

Caryocar costaricensis is listed in Appendix 11 <strong>of</strong> CITES (Wellner and Dickey, 1991).<br />

Pau Rosa (Aniba Duckei) reaches up to 30 m in height. Its wood is used in cabinetry, but its primary<br />

economic importance is the production <strong>of</strong> aromatic essential oil. Pau rosa is locally extinct in many<br />

parts <strong>of</strong> the Amazon, and is currently only found in inaccessible areas. To harvest the essential oil, the<br />

tree is cut, the wood cut into small chips and the wood pulverized. One ton <strong>of</strong> wood chips produces<br />

only 9 kg <strong>of</strong> the oil. (Alencar and Fernandez, 1978). Pau rosa grows well in both full sun and partial<br />

shade. It would be an ideal component in a natural management system, as it is a high value, low<br />

volume product that can be processed locally (Richards, 1993).<br />

Ucuuba (Virola surinamensis) can attain a height <strong>of</strong> 35 m and dbh <strong>of</strong> 45 cm in the forest. The wood is<br />

easy to work and is widely used in light carpentry, for shipping boxes, match sticks, plywood and pulp<br />

for paper. It is heavily exploited for plywood, and commercial timber size trees in the Brazilian<br />

Amazon are extremely rare today. The seeds contain oils useful to the perfume and cosmetics industry,<br />

and are locally used to treat rheumatism, stomach aches and dyspepsia. Cooked bark is used to sterilize<br />

wounds and aid healing. Its sap is used to treat haemorrhoids, and the bark is ground and smoked for<br />

its hallucinogenic effects by many Amazonian tribes (Sampaio in: Clay and Clement, 1993). In the<br />

lower Amazon River Basin, it occurs in association with the buriti (Mauritius flexuosa), assai (Euterpe<br />

oleracea) and ubucu (Manicaria sacci/era) palms, all <strong>of</strong> which have important non-timber uses 56<br />

(Peters et al., 1989; Anderson et al., 1991). These forests should be carefully managed to maintain the<br />

integrity <strong>of</strong> these valuable non-timber product yielding ecosystems. .<br />

Cumaru (Coumarouna odorata) is a large tree <strong>of</strong> the primary forest, attaining up to 30 m, and widely<br />

distributed throughout the Neotropics. Its primary value is its very heavy timber which is used in naval<br />

construction, for truck and train wagons and for high-quality cabinetry. Historically, the extraction <strong>of</strong><br />

coumarin from cumaru seeds was nearly as important as its timber. This perfumed oil was used in the<br />

perfume and cosmetic industries and to flavour tobacco. It is suggested by some authors that markets<br />

for coumarin could increase in the near future due to rising interest in alternative sources <strong>of</strong> vegetable<br />

oils for personal care products (Oay and Clement, 1993). Water extracted from the bark is used as an<br />

antispasmodic and general tonic, and the seeds are occasionally used to make ornamental necklaces and<br />

other handicrafts. Cumaru can grow in both full sun and partial shade, making it a good option for<br />

reforestation or agr<strong>of</strong>orestry. (Sampaio in: Oay and Clement, 1993).<br />

S6 Buriti (Mauritius flexuosa) produces one <strong>of</strong> the most important market fruits in western Amazonian; the thin, oily<br />

mesocarp is eaten raw, or processed into a sweet paste for candies, beverages, and ice creams; the mesocarp and seed<br />

kemal also yield good quality oil. Local inhabitants tend to locate plantations <strong>of</strong> cacao (Theobroma cacao) at the base<br />

<strong>of</strong> these palms (peters et al., 1989; Anderson et al., 1991).<br />

Assai (Euterpe oleracea) is an important source <strong>of</strong> palm heart in Brazil, and its fruit pulp provides a beverage that is a<br />

staple component <strong>of</strong> the regional diet. Local inhabitants actively manipulate the floodplain forests to promote the fruit<br />

<strong>of</strong> this palm (peters et al., 1989; Anderson et al., 1989).


Bacuri (Platonia esculenta) is a mid-to-upper canopy species, requiring full sun for good growth and<br />

yield. It is one <strong>of</strong> the most popular fruits in the Belem market and is used in jams, ice creams and as is.<br />

The yellow latex (both from the fruit and the trunk) has medicinal properties, specifically when applied<br />

topically for skin ailments. The heavy wood is employed in general carpentry and furniture making, as<br />

well as in civil and marine construction. Because it does well on nutrient poor soils and in full sunlight,<br />

bacuri could become a useful species for recuperation <strong>of</strong> degraded lands (Clement, 1993).<br />

Jatoba (Hymenaea courbaril) is generally a large tree, attaining 40 m in height and 2 m dbh. It is<br />

occasionally cultivated and frequently managed by local people, for example by being protected when a<br />

swidden clearing is opened. The wood is very heavy and durable and is used in heavy construction,<br />

such as hydraulic work, truck bodies, railroad ties, etc. The resin that exudes from the trunk, the<br />

branches and the fruit pericarp solidifies on the tree or fall in chunks to the ground, at which time it is<br />

collected. This resin is used industrially in the fabrication <strong>of</strong> varnishes, is used locally to make kitchen<br />

utensils, and as a sealant to reduce water penetration. In popular medicine, jatoba bark is used as a<br />

vennifuge and to treat cystitis; the sap, when mixed with bee's honey, is used to treat bronchitis and<br />

various heart ailments (Ferreira and Sampaio in Clay and Clement, 1993).<br />

VIII. Traditional Management <strong>of</strong> Neotropical Forests for Timber and Non­<br />

Timber Products<br />

Traditional forest management has long relied on complex ecological processes to maximize the<br />

harvest <strong>of</strong> a wide variety <strong>of</strong> forest products. Unlike natural forest management geared primarily to<br />

timber production, which usually results in the coarsening <strong>of</strong> forest structure and depletion <strong>of</strong> species<br />

diversity, the traditional management <strong>of</strong> the forest for a multitude <strong>of</strong> products has usually promoted<br />

and required the maintenance <strong>of</strong> forest structural and species diversity (Gomez Pompa and Burley,<br />

1991; Alcorn, 1990; Posey, 1989; Anderson, 1990).<br />

Indigenous peoples have long managed critical forest resources, rather than just "adapting" to them,<br />

and over the millennia have transfonned much <strong>of</strong> the vegetation and forest types in the tropics<br />

through these elaborate systems <strong>of</strong> forest managemenr 7 (Balee, 1989; 1988; Posey, 1992; Schultes,<br />

1992). Many "natural" forests, therefore, may in fact represent arrested successional forest once<br />

managed by people, including the palm, bamboo, liana and Brazil nut forests. While natural forest<br />

management is generally used to mean the focused management <strong>of</strong> a particular forest product ­<br />

timber-producing trees, non-timber products, game, etc. - indigenous systems <strong>of</strong> forest management<br />

focus on processes, not only products, although products result from these processes (Alcorn, 1989).<br />

For example, the Ka'apor Indians <strong>of</strong> the Brazilian Amazon manage the forest to produce vegetational<br />

zones in different phases <strong>of</strong> recovery. Each vegetational zone is managed with varying degrees <strong>of</strong><br />

intensity, decreasing from house gardens to young swidden, old swidden, and fallow. These diverse<br />

vegetational zones create environments in which artificially high densities <strong>of</strong> useful plants and<br />

57 It is suggested that at least 11.8% <strong>of</strong> the terra firme forest in the Brazilian Amazon alone is <strong>of</strong> archaic, cultural<br />

origin (BaBe, 1989).<br />

35


36<br />

animals thrive to produce diverse products 58 (Balee and Gely, 1989). Similarly, the Mexican Husatec<br />

and Peruvian Bora fanners create mosaics <strong>of</strong> eco-units. In addition to the opened lands for<br />

agriculture, at any given time secondary successional species are reproducing somewhere in the<br />

mosaic and mature forest species are reproducing somewhere else. In this way, the elements<br />

necessary to regenerate forests are retained in the system (Alcom, 1990)59.<br />

Traditional agroecosystems focus on processes, rather than spatial structures, the usual subject <strong>of</strong><br />

western attention. They are a fluid complex <strong>of</strong> planted fields, fallows, savanna, dooryards, forests,<br />

rivers, and river banks, rather than distinct, static structures (Alcom, 1989; 1993). According to<br />

Alcom (199), indigenous agr<strong>of</strong>orestry systems share seven attributes: they take advantage <strong>of</strong> the<br />

variety <strong>of</strong> native trees and tree communities, planting or protecting those most desired; rely on natural<br />

successional processes to produce resources, improve and protect soils, and reduce pest problems;<br />

use natural environmental variation; incorporate numerous crop and native species; are flexible and<br />

personalisable, and may vary by site or year; spread risks by retaining diversity; and maintain a<br />

reliable back-up to meet needs should other sources fail.<br />

Indigenous sequential agr<strong>of</strong>orestry systems that integrate secondary successional vegetation are<br />

usually complimented by managed forest grove systems from which fruits, wood, and game are<br />

harvested. In Huastec communities, 25% <strong>of</strong> a community's land is held in forested groves, usually<br />

situated along creeks or on steep slopes where they prevent soil erosion and protect watersheds. They<br />

are managed by a combination <strong>of</strong> arboriculture and natural forest management to create a mixture<br />

<strong>of</strong> native species <strong>of</strong> primary and secondary forest, and introduced species. A typical forest grove<br />

contains over 300 species and 90% have known uses (as food, construction material, medicine,<br />

firewood and livestock forage)60 (Alcom, 1990).<br />

58 Over 76% <strong>of</strong> the plant species used by the Kayapo Indians <strong>of</strong> Brazil are not "domesticated" nor can they be<br />

considered "wild" since they have been systematically selected for desimble traits and propagated in a variety <strong>of</strong> habitats<br />

(posey, 1992).<br />

59 Irvine (1989) found that the Runa Indian community in Ecuador have an important impact on forest structure and<br />

composition through succession management. Regeneration in the fallow forest is managed to develop a characteristic<br />

form, and so alters the plant composition <strong>of</strong> the developing forests. Planting <strong>of</strong> species that can mature during the fallow<br />

period occurs early in the agricultural cycle. Light-demanding fruit trees are planted concurrently with manioc, for<br />

example, to benefit from the larger gap in the forest at this stage. Selective cutting <strong>of</strong> the forest understorey or overstory<br />

during garden clearing, and subsequent weeding, also favours certain forest species as sources <strong>of</strong> regeneration. Both<br />

planting and protection <strong>of</strong> valuable species vary from the careful to the casual. Managed fallows had a more diverse<br />

canopy, with a higher proportion <strong>of</strong>enrichment (8-19% planted species); unmanaged fallows were dominated by a unifonn<br />

canopy <strong>of</strong> Caecropia spp.<br />

60 An ethnobotanical survey conducted by Boom (1989) <strong>of</strong> the plants used by the Chacabo Indians <strong>of</strong> Amazonian<br />

Bolivia showed that 360 (or 82%, <strong>of</strong> the tree species and 95% <strong>of</strong> the individual trees) individuals in one hectare <strong>of</strong> forest<br />

near their principle village, Alto Ivon, were used: food (33 spp., 264 individuals), fuel (14 spp., 163 individuals),<br />

construction and crafts (23 spp., 225 individuals), medicines (23 spp., 271 individuals) and commercial sources <strong>of</strong> cash<br />

(2 species: Hevea brasiliensis and Bertholletia excelsa).


38<br />

With slight modification, this could be a list <strong>of</strong> management activities employed by the forester.<br />

Rarely, however, are these traditional management systems - developed over millennia by the<br />

inhabitants <strong>of</strong> extremely unique and complex ecosystems - seriously considered in the design <strong>of</strong><br />

natural forest management 63 • Elaborate scientific studies are undertaken to articulate the subtleties <strong>of</strong><br />

what <strong>of</strong>ten amounts to a small component <strong>of</strong> traditional management systems, and extensive<br />

ignorance <strong>of</strong> the ecology <strong>of</strong> tropical forests is widely pr<strong>of</strong>essed. While this research is <strong>of</strong> academic<br />

interest, it will take many decades to arrive at a well-articulated scientific understanding <strong>of</strong> even a<br />

small portion <strong>of</strong> the tropics, and the forest manager must rely on more immediate guidance. For lowintensity,<br />

multi-purpose natural forest management, traditional management systems provide an<br />

enormous headstart 64 (Alcorn, 1990; Gomez-Pompa, 1991; Salick, 1992; Posey, 1988; Zemer, 1993;<br />

Lamb, 1991).<br />

IX. Natural Forest Management and the Conservation <strong>of</strong> Biodiversity<br />

Approximately 1.4,000,000 species <strong>of</strong> animals and plants are known to science, and it is thought that<br />

more than three times this number exist (Sayer, 1991). Tropical forests, which cover only 7% <strong>of</strong> the<br />

earth's land mass, are estimated to contain well over half its species. One ha <strong>of</strong> tropical forest is<br />

estimated to contain between 100 and 300 trees species alone, with most species being represented<br />

by only one or two individuals per hectare. The general trend <strong>of</strong> high species diversity with low<br />

species density has been documented repeatedly in tropical forest inventories (Peters, 1993). As a<br />

result, tropical forest species appear especially prone to extinctions and sustainable resource<br />

63 Admittedly, there are numerous obstacles to this, not least <strong>of</strong> which is the lack <strong>of</strong> respect western-style managers<br />

<strong>of</strong>ten have for traditional systems <strong>of</strong> management that are based not upon fonnal science but rather on customary practice,<br />

cultural traditions, and local knowledge <strong>of</strong> land and natural resources (Othole and Anton, 1993; Poole, P. 1993; Berkes<br />

et al., 1991). Traditional, largely decentralized and consensus-based systems <strong>of</strong> management that are enforced through<br />

social sanctions, contrast markedly with the "western" hierarchical, centralized, regulatory, and time-based systems <strong>of</strong><br />

management.<br />

As the Zuni Indians Othole and Anyon (1993) state: "because the entire legal, regulatory, and guideline framework is a<br />

non-Indian construct, it is <strong>of</strong>ten difficult to fit the needs <strong>of</strong> the developer and agency with the needs <strong>of</strong> the tribe ... even<br />

the word 'property' albeit a necessity <strong>of</strong> tenninology because <strong>of</strong> the language <strong>of</strong> federal law ... can raise serious concerns<br />

within the traditional and religious leadership".<br />

64 For example, Salick (1992) has found that the increased dominance <strong>of</strong> secondary forest species and the importance<br />

<strong>of</strong> coppicing as a source <strong>of</strong> regeneration in the strip-clear cutting silvicultural system employed in Palcazu Peru,<br />

resembled fallows common to indigenous systems <strong>of</strong> forest management from the area. She recommended using<br />

indigenous forest and fallow management as the theoretical underpinning to natural forest management; by applying these<br />

techniques, forests may be more productive. For example, to maximize regeneration <strong>of</strong> desired species in the relatively<br />

large gaps <strong>of</strong> the strip clear-cut, in addition to managing seed sources, emphasis should be placed on management<br />

practices that encourage healthy coppicing, as has been done successfully in the indigenous system (Salick, 1992).<br />

Phillips and Gentry (1993) have developed a simple quantitative technique for evaluating the relative usefulness <strong>of</strong> plants<br />

to people; they feel that compilation-style approaches to ethnobotany must be modified by developing methods that allow<br />

researchers to quantitatively describe and analyze patterns in what they study. In this way, ethnobotanical studies will<br />

provide scientifically higher-quality infonnation on why and how people use plants, and so might more easily be<br />

incorpomted into the research and design <strong>of</strong> management systems for natural forests by other scientists, including<br />

foresters.


exploitation is a difficult management problem (Sayer, 1991; Peters, 1993)65. As A.D. John (1992)<br />

stated, "in practice, the precise environmental factors to which species may be responding are difficult<br />

to define". As a result, we cannot accurately determine the impact <strong>of</strong> logging on biodiversity and the<br />

many forest products that draw on this diversity. The implications <strong>of</strong> our ignorance <strong>of</strong> species<br />

diversity and tropical ecosystems for natural forest management are clear - we know little, are only<br />

very slowly learning more, and so must err on the side <strong>of</strong> caution.<br />

Nepsted et al. (1992) describe two types <strong>of</strong> biotic impoverishment, or ecological degradation, that<br />

result from human use <strong>of</strong> forest resources: population impoverishment and ecosystem<br />

impoverishment 66 . Ecosystem impoverishment always influences populations, but we do not have<br />

sufficient information to predict the exact nature <strong>of</strong> these impacts. The harvest <strong>of</strong> non-timber forest<br />

products, for example <strong>of</strong> tapir and brazil nuts in the extractive reserve <strong>of</strong> Porongaba in the Amazon,<br />

could result in population impoverishment, but rarely ecosystem impoverishment. Logging in the<br />

same area can lead to the population impoverishment <strong>of</strong> roughly 100 tree species and an unknown<br />

number <strong>of</strong> animal species, and potentially severe ecosystem impoverishment due largely to the 50%<br />

canopy reduction with which logging is associated (Nepstad et aI., 1989).<br />

Gaps in the natural forest created from disturbances such as landslides, earthquakes, rivers switching<br />

course, or individual tree falls are theorized to contribute to species diversity, but gaps created by<br />

logging regimes generally do not result in more diverse forests as a whole (Brown, 1992). In general,<br />

the smaller the scale <strong>of</strong> disturbance, the more likely local structural, floristic and faunistic diversity<br />

will be temporarily enhanced or remain the same. Larger scale disturbances tend to simplify<br />

ecosystems, which results in a loss <strong>of</strong> biodiversity both locally and regionally. There is always a<br />

trade-<strong>of</strong>f, then, between the intensity <strong>of</strong> timber production and the conservation <strong>of</strong> biodiversity (ITTO,<br />

Draft Guidelines for Biological Diversity Conservation, 1992; Poore and Sayer, 1991).<br />

65 There are exceptions to the rule <strong>of</strong> high species diversity in tropical forests. Oligarchic forests, dominated by only<br />

one or two tree species, occupy tens <strong>of</strong> millions <strong>of</strong> hectares in Amazonia. In many cases, the dominant species produce<br />

fruits, seeds or oils <strong>of</strong> economic importance. In terms <strong>of</strong> density and yield, oligarchic forests may rival many <strong>of</strong> the<br />

commercial fruit plantations in the tropics, and so are a valuable source <strong>of</strong> non-timber forest products (peters, 1993; Peters<br />

et al., 1989). For example, camu-camu (Myrciaria dubia), a shrub or small tree native to the floodplains <strong>of</strong> Amazonia,<br />

is <strong>of</strong>ten found in large, mono-specific populations; isolated individuals are rarely encountered. In natum! populations,<br />

12,310 individuals/hectare have been found (C. Peters, pers. comm, in Clay and Clement, 1993). Camu-camu fruits<br />

contain one <strong>of</strong> the highest concentrations <strong>of</strong> vitamin C in the plant kingdom (2,000-2,994 mg ascorbic acid/lOO g <strong>of</strong> fruit<br />

pulp), and are widely used locally in juices, ice creams and liqueurs (peters et al., 1989).<br />

66 Population impoverishment results when the genetic diversity or abundance <strong>of</strong> a population declines as a result <strong>of</strong><br />

human activity. Hunting or harvesting practices that favour certain types <strong>of</strong> animal or plant characteristics or that exceed<br />

the regenerative capacity <strong>of</strong> the population impoverish the population genetically and structurally. Populations can also<br />

decline as the indirect result <strong>of</strong> human activity if pollinators or seed dispersal agents are eliminated. Ecosystem<br />

impoverishment results when forest use alters the structural and functional integrity <strong>of</strong> the forest, changing its ability to<br />

regulate the storage and flow <strong>of</strong> water, energy, carbon, and mineral nutrients (Nepsted et al., 1992).<br />

39


40<br />

Timber harvesting may not always lead to a reduction in species numbers, but can produce qualitative<br />

changes whereby generalists replace old-growth specialists and a few species come to numerically<br />

dominate 67 (IITO, 1992; John, A.D., 1992; Dahaban, Nordin and Bennett, 1992; Kemp et al., 1993).<br />

Logging damage is essentially random, in that it is spread over all tree taxa. As a result, rare species<br />

are particularly susceptible to the impacts <strong>of</strong> logging, especially those that are also highly valued<br />

timbers 68 (John, A.D., 1992). Timber species with a limited geographic range, poor dispersal ability,<br />

and few seedlings and saplings in the understorey are generally less equipped to deal with logging<br />

pressures than those widely distributed, with long seed dispersal distances and with numerous<br />

seedlings and saplings in the understorey. Amazonian species such as Swietenia macrophylla,<br />

Vouacapoua americana, Torresia acreana (used locally for medicine and perfume) and Euxylophora<br />

paraensis were found to be susceptible to pressures in the face <strong>of</strong> intensive logging, while Tabebuia<br />

serratifolia, Hura creptans (exudate used locally for poisons) and Astronium urundeuva (wood used<br />

locally for crafts) might be favoured by changes induced by logging. Most species fell between the<br />

two extremes 69 (Martini et al., 1992). The qualitative changes in forest composition as a result <strong>of</strong><br />

logging can promote some timber species used locally for NTFPs, while reducing populations <strong>of</strong><br />

others. But most communities rely on a variety and range <strong>of</strong> forest products to meet their needs and<br />

a reduction in the diversity <strong>of</strong> available species can have significant repercussions on local<br />

consumption and trading patterns. Similarly, the conservation <strong>of</strong> biodiversity does not depend upon<br />

the total number <strong>of</strong> species in a small area, but the number within an entire region. While the<br />

disturbances created by logging, differential species success in varying gap sizes, and the resulting<br />

67 For example, in preliminary studies in unlogged and logged dipterocarp forest in Sarawak, environmental changes<br />

have produced considerable drops in termite species richness, favouring mound-building species over those constructing<br />

simple nests in soil (John, A.D., 1992). Another study in dipterocarp forest in Sarawak demonstrated that following<br />

logging the trend is for edge or colonizer species to replace primary forest species. A number <strong>of</strong> species appeared that<br />

had never been seen before in the forest, but are common to secondary forests and gardens, such as the plain-tail squirrel<br />

(Callosciurus notatus) and shrews (Tupaia sp.). Those that disappeared after logging were all endangered or totally<br />

protected species or both, such as the clouded leopard (Ne<strong>of</strong>elis nebulosa), sun bear (Helarctos malayanus) and oriental<br />

small-clawed otter (Aonyx conerea) (Dahaban, Nordin, and Bennett, 1992). Stoekdale (pers. comm., 1993) has found that<br />

in 12 year old logged forest in East Kalimantan, populations <strong>of</strong> the slow-growing, high quality Calamus and Daemonorops<br />

rattan species have been reduced, while the lower value, pioneer-like Korthalsia species have increased in number. This<br />

will obviously have important economic implications for future harvests <strong>of</strong> rattan in the region.<br />

68 For example, Brazilwood (Caesalpinia echinata) has been eradicated from most <strong>of</strong> its former range in Amazonian<br />

(John, A.D., 1992). Ceiba pentandra, which has religious importance for many traditional communities, and the seeds<br />

<strong>of</strong> which are used locally to stuff mattresses and the wood for construction, was extirpated from Iquitos, Peru by intensive<br />

logging operations (Gentry and Vasquez, 1988; Orejuela, 1992).<br />

69 The eight parameters used to evaluate a species' ability to resist the negative impacts <strong>of</strong> logging include: 1)<br />

effective long-distance dispersal ability; 2) abundance <strong>of</strong> saplings in the forest regeneration; 3) rapid growth in full sun;<br />

4) ability to resprout; 5) capacity to withstand fire (by virtue <strong>of</strong> thick bark); 6) broad geographic distribution; 7) even<br />

distribution over the landscape; and 8) high density <strong>of</strong> adults (Martini et al., 1992).


increases in "edges" can produce localized increases in diversity, it does not increase biodiversity for<br />

the region as a whole 70 (Dahaban, Nordin and Bennett, 1992).<br />

Logging can also impact the survival or density <strong>of</strong> keystone or pivotal species that control the<br />

structure <strong>of</strong> the community and help determine which other species are present (WRI et al., 1992).<br />

At Kutai, East Kalimantan, for example, it was found that the "keystone" Ficus species are<br />

significantly associated with some <strong>of</strong> the major timber species which logging selectively depletes<br />

(Whitmore, 1991)71. In these cases, population impoverishment, as described by Nepstad et al.<br />

(1992), can lead to ecosystem impoverishment, because keystone species are not "redundant" - there<br />

do not exist several other species that perform similar ecosystem functions.<br />

When species are left in small isolated populations they become susceptible to extinction by random<br />

events and genetic deterioration. This, again, is a potentially hidden side-effect <strong>of</strong> logging operations,<br />

and is particularly important in tropical regions where local endemism is high (John, A.D., 1992).<br />

Heywood and Stuart (1992) state that the loss <strong>of</strong> genetic variation in species' populations is at least<br />

as important as the loss <strong>of</strong> entire species, and is the "essence <strong>of</strong> the process <strong>of</strong> extinction". Genetic<br />

variation, like soil fertility, can be impossible or extremely slow to rehabilitate through management<br />

schemes. Therefore, fragmentation <strong>of</strong> forests during logging operations should be kept to a minimum;<br />

logging should be staggered so that various areas are at various stages <strong>of</strong> succession following<br />

disturbance and mature stands lie in close proximity to each other (WRI et al., 1992). Uncontrolled<br />

selective logging can also eliminate the more desirable genetic characteristics <strong>of</strong> a timber species by<br />

systematically "creaming" or removing individuals that display valuable traits (Uhl, 1989; Oldfield,<br />

1989; Schmidt, 1991; Clay and Clement, 1993). The intensive collection <strong>of</strong> fruits from trees that<br />

produce those <strong>of</strong> the highest quality can similarly result in a population dominated by trees <strong>of</strong><br />

marginal economic value (Peters, 1993).<br />

There exists no objective, long-term data on complete logging cycles, so we can only theorize on<br />

impacts over time. Different sites have a wide variation in species abundances and distribution over<br />

even quite small areas, so each forest area must be individually assessed with regards to local species<br />

diversity, endemism and degrees to which logging disturbance produces negative effects. In the short<br />

tenn, studies <strong>of</strong>ten demonstrate little decline in species diversity, so long-term studies must be<br />

implemented that detect not only changes in species numbers but in density, as well. The study <strong>of</strong><br />

local structural, floral and faunal diversity can be incorporated into permanent sample plots set up<br />

to measure the growth, mortality, and regeneration rates <strong>of</strong> commercial timber (and non-timber)<br />

species, and other changes in the forest over time (Alder and Synnott, 1992).<br />

70 There is little available information on the effect <strong>of</strong> disturbance on most species, nor on the relative effects <strong>of</strong><br />

shorter and longer felling cycles. (Ashton, pers. comm. in Heywood and Stuan, 1992). Species loss is a logarithmic<br />

function, and so repeated logging, or re-logging before full regeneration <strong>of</strong> the forest has occurred could lead to severe<br />

and permanent effects on biodiversity (John, A.D., 1992) Many studies which have attempted to detennine the impact<br />

<strong>of</strong> selective logging or other disturbances on species diversity have produced inconclusive results due to the variety <strong>of</strong><br />

factors involved (John, R.D., 1992).<br />

For example, RJ. John (1992) found that in Papua New Guinea he could not adequately detennine the effects <strong>of</strong> selective<br />

logging on species diversity. He suggests that the major impacts <strong>of</strong> logging are the physical effects <strong>of</strong> the removal <strong>of</strong> logs<br />

and post-logging development <strong>of</strong> lands made accessible through logging operations.<br />

71 Many <strong>of</strong> these species are used extensively by local peoples, as well. In the Gunung Leuser National Park, for<br />

example, Ficus species are collected for traditional medical treatments (Sayer, 1991).<br />

41


42<br />

The management <strong>of</strong> natural tropical forests for timber production will inevitably result in a loss <strong>of</strong><br />

species and structural diversity, but these forests can form an important component <strong>of</strong> national landuse<br />

systems that attempt to conserve biodiversity. In no tropical country today will the lands<br />

demarcated as strict nature reserves protect an adequate proportion <strong>of</strong> that country's biodiversity<br />

(Hansen et al., 1991; Blockhus et al., 1992; WRI et aI., 1992). Forest management for non-timber<br />

products alone is usually more compatible with biodiversity conservation 72 (Salick, 1992; Gomez­<br />

Pompa, 1990). But because a significant proportion <strong>of</strong> the world's remaining natural forests are<br />

allocated for timber production, the survival <strong>of</strong> biodiversity depends in large part on the<br />

implementation <strong>of</strong> natural forest management systems that integrate timber, non-timber products and<br />

conservation objectives.<br />

Because forests are dynamic, they are reasonably robust in their ability to recover from localized and<br />

periodic disturbance. The selective removal <strong>of</strong> a small volume <strong>of</strong> timber and non-timber forest<br />

products, followed by management that allows regeneration to occur, can retain a significant portion<br />

<strong>of</strong> species. The number <strong>of</strong> species whose populations are critically reduced by logging, and so the<br />

speed with which the original animal and plant communities are reestablished, is related to: the<br />

degree to which the forest is disturbed; the proximity <strong>of</strong> undisturbed patches which can act as refuges<br />

for mobile species; and the maintenance or return <strong>of</strong> substantial portions <strong>of</strong> the forest to the mature<br />

phase structural complexity in which dominance by a few species becomes impossible and diversity<br />

is maintained (IITO, 1992; Reid, 1992; Blockhus et al., 1992; Brown and Brown, 1992, John, A.D.,<br />

1992).<br />

The Impact <strong>of</strong> Natural Forest Management on Wildlife<br />

Animal-plant relationships are a significant factor in the long-term success <strong>of</strong> natural forest<br />

management. Regeneration <strong>of</strong> valuable commercial timber species does not take place in an<br />

ecological vacuum, and is not based solely on the size and frequency with which gaps are opened<br />

during logging operations and the effectiveness <strong>of</strong> silvicultural treatments. Through predation,<br />

pollination, and seed dispersal, animals play an integral role the regeneration and survival <strong>of</strong> tree<br />

species.<br />

There exist wide variations in the response <strong>of</strong> wildlife to changed conditions <strong>of</strong> habitat and food<br />

supply due to logging. Small, less mobile species are most likely to suffer population reductions as<br />

a result <strong>of</strong> patchy food sources, and this may affect ranging patterns, breeding success and even gene<br />

flow. Large-bodied, wide-ranging species with generalized diets that may opt for alternative sources<br />

<strong>of</strong> food will fare best (John, A.D., 1992). Whitmore notes that elephants, rhinoceros, pigs, and other<br />

species that feed generally on lush herbs and shrubs found along valleys and on landslips actually<br />

benefit from the young regrowth along logging tracks, while specialists such as cuckoos, trogans and<br />

many insectivores do not fair as well (Whitmore, 1991). But Dahaban, Nordin and Bennett (1992)<br />

found that in dipterocarp forests in Sarawak the diversity and density <strong>of</strong>primates, squirrels, and other<br />

72 Salick (1992) found that forest under more or less traditional management, although containing fewer species than<br />

primary forest, had higher densities <strong>of</strong> those species <strong>of</strong> larger non-timber value, including ipecac and wicker. Selectively<br />

logged forest had fewer species overall and fewer useful non-timber species. Boom (1989) found that the Chacabo Indians<br />

in Bolivia similarly managed forest to increase the proportion <strong>of</strong> useful species in proximity to their village. High species<br />

dominance is <strong>of</strong>ten a reflection <strong>of</strong> past management activities, including the equivalent <strong>of</strong> forest gardens, and the<br />

protection and enrichment <strong>of</strong> valuable species (Gomez-Pompa, 1990).


mammals was severely affected following logging 73 . A.D. John (1991) found in Western Amazonian<br />

that avifaunas in disturbed habitats were dissimilar to that <strong>of</strong> undisturbed forest to a degree<br />

proportional to the extent <strong>of</strong> disturbance. Understorey insectivorous birds with specialized foraging<br />

strategies tend to be least able to adjust to conditions <strong>of</strong> disturbance (John, 1991; Alejandro Grajal,<br />

Wildlife Conservation Society, pers. comm., 1993).<br />

In most tropical countries over 70% <strong>of</strong> resident vertebrate species are commonly reported to be<br />

dependant upon closed forest, and the figure for invertebrates is probably a great deal higher (John,<br />

A.D., 1992). Invertebrates are far more species-rich than vertebrates and have high microhabitat<br />

specialization, so are likely to be severely affected by logging disturbance, but virtually nothing is<br />

known about them (John, A.D., 1992).<br />

Animals in recently logged forest may face three problems concerning food source trees: fewer trees,<br />

different spatial distribution <strong>of</strong> trees, and different patterns <strong>of</strong> fruit and leaf production. The initial<br />

loss during logging <strong>of</strong> trees used as food sources by wildlife might be buffered by increased levels<br />

<strong>of</strong> fruiting and new leaf production stimulated by increased light from the opening <strong>of</strong> the canopy, but<br />

the full range <strong>of</strong> food sources will not return until the regeneration <strong>of</strong> primary forest seedlings reach<br />

reproductive maturity (John, A.D., 1992; 1988)74.<br />

Crome (1991) found that while selectively logged forests can provide an excellent habitat for many<br />

species <strong>of</strong> wildlife, it can threaten specialist frugivorous that are important components <strong>of</strong> the forest<br />

ecosystem and through dispersal impact the demography <strong>of</strong> plant species 75 . Uhl (1989) found that<br />

in the Amazon many <strong>of</strong> the most sought-after timber species produce fruits that are important food<br />

for forest animals. For example, Manilkara huberi, which accounted for 30% <strong>of</strong> the 1988 timber<br />

volume harvested in Para, has cherry-sized fruits that are consumed by parrots, monkeys, coati, deer<br />

and turtles. A.D. John (1988) found in a West Malaysian dipterocarp forest that residual damage from<br />

logging operations drastically reduced overall availability <strong>of</strong> food sources for frugivorous and<br />

folivores, even where timber trees are not themselves used by animals.<br />

Post-logging silvicultural treatments can also negatively impact wildlife populations. By promoting<br />

the regeneration <strong>of</strong> larger proportions <strong>of</strong> commercial species, poisoning the undesirable, and<br />

"enriching" the forest - <strong>of</strong>ten with exotics - silvicultural interventions can alter plant species<br />

composition and destroy species that maintain wildlife populations (Caldecott, 1988; Crome, 1991).<br />

In a study <strong>of</strong> the commercial species <strong>of</strong> preference in Queensland Crome (1991) found that the vast<br />

73 Borean gibbons (Hylobates muellerz) decreased by more than 70% in group density after logging. The group density<br />

<strong>of</strong> maroon langurs (Presbytis rubicunda) dropped about 45% after logging, while that <strong>of</strong> the white-fronted langurs (P.<br />

frontata) decreased by 35%. Giant squirrels (Ratufa affinis) showed a reduction <strong>of</strong> 71 % in individual density after logging.<br />

The diversity <strong>of</strong> species in these areas also decreased markedly. (Dahaban, Nordin, and Bennett, 1992).<br />

74 The uncontrolled collection <strong>of</strong> non-timber forest product reproductive propagules (fruits, nuts/seeds, and oilseeds)<br />

can also disturb food availability for wildlife. In effect, collectors compete with forest frugivorous, reducing the total<br />

supply <strong>of</strong> food resources available to ground-foraging animals. Frugivorous might subsequently migrate <strong>of</strong>f-site and<br />

disrupt seed dispersal and seedling establishment for those species whose seeds require scarification by animals to<br />

germinate (peters, 1993).<br />

75 "Legitimate" frugivorous, which digest only the pericarp and void the seeds intact, require fruits <strong>of</strong> high nutritive<br />

value, are highly specialized and are maintained by a succession <strong>of</strong> appropriate plant species fruiting sequentially. In lean<br />

periods, they rely on "keystone" species and are highly vulnerable to the loss <strong>of</strong> a food source (Crome, 1991).<br />

43


44<br />

majority <strong>of</strong> frugivorous rely on uncommercial species, rather than the increasing proportions <strong>of</strong> highvalue<br />

commercial species with woody fruits and wind dispersed seeds. Silvicultural regimes,<br />

therefore, promoted a decrease in the populations <strong>of</strong> trees needed during lean times for food.<br />

Enrichment planting <strong>of</strong> local wild seedlings, such as those recommended by the ITTO, may not<br />

always adequately protect wildlife diversity, although they will undoubtably be an improvement on<br />

planting <strong>of</strong> exotics 76 (ITTO, Draft Guidelines for Biological Diversity Conservation, 1992). In<br />

Venezuela, enrichment plantings <strong>of</strong> non-indigenous commercial species in strips were found to<br />

contain 30% fewer species than the surrounding forest; this was less than the number in the logged<br />

over forest left to naturally regenerate (Alejandro Grajal, Wildlife Conservation Society, pers. comm.,<br />

1993).<br />

In Sarawak there is typically a sharp decline in wildlife harvests within ten years <strong>of</strong> an area first<br />

being logged due to a combination <strong>of</strong> ecological, social and economic factors 77 : the immediate<br />

physical disturbance resulting from logging; the long-term change in the forest composition and<br />

structure; the influx <strong>of</strong> timber company workers, many <strong>of</strong> whom will hunt themselves 78 ; and<br />

improved lines <strong>of</strong> communication and <strong>of</strong>ten additional cash incomes for many rural people. These<br />

factors result in increased hunting pressure as people take advantage <strong>of</strong> new markets and the access<br />

provided by logging road openings into previously undisturbed forest (Caldecott, 1989; John, R.J.,<br />

1992; Dahaban, Nordin and Bennett, 1992; Wilkie et aI., 1992). Some species are particularly<br />

vulnerable to extermination by hunting, such as rhinoceros, gibbons, bears, leaf monkeys, proboscis<br />

monkeys, clouded leopards and marine turtles (Caldecott, 1988; Dahaban, Nordin and Bennett, 1992).<br />

In this way, the increased opportunities for feeding on herbs along logging tracks suggested by<br />

Whitmore, and the resilience John (1992) suggests primates might show following logging operations,<br />

are <strong>of</strong>f-set by increased hunting pressures (John, A.D., 1986; Redford, 1991; Whitmore, 1991).<br />

In Sarawak, Dahaban, Nordin and Bennett (1992) found that the access provided by logging roads<br />

resulted in a sharp increase in hunting pressures on once remote forest (about 40 km from the nearest<br />

human settlement). Hunting was <strong>of</strong>ten carried out from vehicles, and hunters used spotlights at night<br />

to stun animals. In the Republic <strong>of</strong> Congo, Wilkie et al. (1992) found that the highly selective<br />

logging practised in the concession under study was unlikely to affect faunal populations, but that<br />

the active and passive facilitation providing by logging operations (vehicles, roads, facilities and<br />

company time were <strong>of</strong>ten used in support <strong>of</strong> poaching) increased pressures on wildlife substantially.<br />

76 Although F. Omari (pers. comm. in John, A.D., 1992) found that the impact on food sources for local wildlife was<br />

minimized by the replanting <strong>of</strong> Calophyllum inophyllum on Zanzibar Island.<br />

77 Hunting in logged areas results in a sharp decline in the estimated wild meat ration per person/year from 54 to 18<br />

g. This effect is exacerbated by a serious decline in riverine fish stocks due to mud and diesel pollution from logging<br />

operations (Caldecott, 1988).<br />

78 In the Neotropics as well, hunting is frequently perfonned by people engaged in forest extraction activities who<br />

are given ammunition but little food, and are expected to hunt to feed themselves. Venezuela et al. (1988) found that in<br />

a watershed near Iquitos, Peru, lumbennen accounted for 51 % <strong>of</strong>the ungulate harvest, illegal commercial hunters for 11 %<br />

and subsistence hunters for only 38% (Redford, 1991).


Many <strong>of</strong> the most popular game animals are also those which Terborgh refers to as having a<br />

"stabilizing function" and through "indirect effects" maintain the diversity <strong>of</strong> tropical forests 79<br />

(Terborgh, 1988). In Mexico, Dirzo and Miranda (1990 in Redford) compared two tropical forests,<br />

one with its full compliment <strong>of</strong> large mammals (peccaries, deer and tapir) and the other in which<br />

these species had been extirpated by hunters. The hunted forest was typified by seedling carpets, piles<br />

<strong>of</strong> uneaten rotting fruits and seeds, and herbs and seedlings undamaged by mammalian herbivores<br />

(Redford, 1991). Although many important forest animals may not yet be extinct, their populations<br />

may have been reduced to such an extent that they no longer perform their ecological functions SO<br />

(Redford, 1991). A decrease in the number <strong>of</strong> animals in a forest area can also lead to a number <strong>of</strong><br />

more subtle, largely unrecorded, changes within populations. These include changed reproductive<br />

behaviour, sex and age ratios, average body size, patterns in habitat use, and the genetic structure <strong>of</strong><br />

a population 81 •<br />

Unfortunately, the pressure to log primary tropical forests is too great to be able to wait for the type<br />

<strong>of</strong> information needed to determine the ecological mechanisms by which sustainable natural forest<br />

management is possible. In the interim, a compromise must be practised in which indicator species,<br />

coupled with long-term monitoring, allows us to determine a maximum threshold for damage to<br />

wildlife and biodiversity resulting from timber production.<br />

x. Conclusion<br />

In many ways the integration <strong>of</strong> non-timber and timber products is a new arena <strong>of</strong> forest<br />

management. Although forest-dwelling peoples have harvested both timber and NTFPs for millennia,<br />

and the harvest <strong>of</strong> NTFPs <strong>of</strong>ten occurs vicariously in recently logged forest or forests managed for<br />

timber production, the industrial, usually capital-intensive forestry practised over the past half a<br />

century, has been almost entirely divorced from non-timber concerns. There has also existed the<br />

perception that what isn't marketed in scale isn't important, and isn't "developed" (Redford, 1991;<br />

FAO, 1991).<br />

As a result, parallel fields <strong>of</strong> study, literature and academic departments have developed for the same<br />

forests. Ethnobotanists, anthropologists, "social" foresters, and cultural ecologists have a welldeveloped<br />

literature and discussion underway on the role <strong>of</strong> NTFPs in local forest management.<br />

Foresters, on the other hand, have developed their own plans for the same forests. Rarely will these<br />

pr<strong>of</strong>essionals interact - rarely will they read the same literature, attend the same conferences, or<br />

collaborate on projects. Lip service is being paid to the integration <strong>of</strong> the forest sciences and "multi-<br />

79 ttIndirect effects" refer to "the propagation <strong>of</strong> perturbations through one or more trophic levels in an ecosystem,<br />

so that consequences are felt in organisms that may seem far removed, both ecologically and taxonomically, from the<br />

subjects <strong>of</strong> perturbation" (Terborgh, 1988).<br />

80 "Ecological extinction tt is defined as "the reduction <strong>of</strong> a species to such a low abundance that although it is still<br />

present in the community it no longer interacts significantly with other species" (Estes et al., 1989).<br />

81 A.D. Johns (1992) describes changes in hummingbird communities in the neotropics as a result <strong>of</strong> logging.<br />

Hummingbirds are generally co-adapted with communities <strong>of</strong> flowering plants, with the different species <strong>of</strong> hummingbird<br />

diverging in bill shape to exploit corolla design. In the regeneration <strong>of</strong> logged forest, much available nectar comes from<br />

colonizing shrubs and climbers with flowers <strong>of</strong> generalized corolla shape. This resulted in the social reorganization <strong>of</strong><br />

hummingbird communities to one <strong>of</strong> aggressive defense <strong>of</strong> resources (interference competition).<br />

45


46<br />

disciplinarytf approaches to forest management, but schemes to integrate non-timber forest products<br />

and multiple-use with timber production are virtually non-existent. As a result, the successful<br />

practical integration <strong>of</strong> timber and non-timber forest products into natural management <strong>of</strong> tropical<br />

forests may still be some way <strong>of</strong>f.<br />

In this context, the certification and labelling <strong>of</strong> tropical timbers from sustainable sources has<br />

emerged. In their efforts to achieve "socially beneficial, economically viable, and environmentally<br />

benign" timber extraction, certifiers and their potential accreditors in the form <strong>of</strong> the Forest<br />

Stewardship Council, have come to realize that non-timber forest products are an integral component<br />

<strong>of</strong> sustainable natural forest management (Forest Stewardship, Draft Principles and Criteria, 1993;<br />

Donovan, 1993). As a result, they have made a significant step towards bridging the gap that lies<br />

between the non-timber product sphere <strong>of</strong> expertise and the timber. In the Draft Principles and<br />

Criteria <strong>of</strong> the Forest Stewardship Council, for example, "non-timber forest resources must be<br />

inventoried and continued access to such resources by local people incorporated into forest<br />

management" and "research must be conducted on the full range <strong>of</strong> timber and non-timber forest<br />

products and services found in the forest area, and incorporated into management planning" (FSC,<br />

1993).<br />

Following are some basic suggestions relevant to the integration <strong>of</strong> non-timber product and timber<br />

forest management. General principles <strong>of</strong> sound forest management and social and environmental<br />

responsibility, as outlined in the Forest Stewardship Council's Principles and Criteria, the ITIO<br />

Guidelines for the Conservation <strong>of</strong> Biological Diversity in Forests Managed for Timber and<br />

Guidelinesfor Sustainable Management, and The Smart Wood Certification Program Guidelines, will<br />

not be reiterated here.<br />

It is important that foresters, timber companies and governments do not try to implement integrated<br />

forest management on their own. Collaborations with local, national and international researchers,<br />

NGOs, and development agencies, can bring to the operation a knowledge <strong>of</strong> local forests, species,<br />

NTFPs, and people that will compliment that <strong>of</strong> the timber specialists.<br />

NTFPs should not be an add-on, randomly attached to established management plans. Inventories<br />

<strong>of</strong> species, assessments <strong>of</strong> their social and economic value to local people, and a rough estimate<br />

<strong>of</strong> existing trade patterns, must be quantified in some way prior to the setting <strong>of</strong> management<br />

objectives (De beer and McDermott (1989), for example, provide a list <strong>of</strong> quantifiable indicators<br />

<strong>of</strong> the economic value <strong>of</strong> NTFPs to rural people).<br />

Similarly, local people should not be "consulted tf regarding an existing management plan, but<br />

should be integrated into the process by which priorities and objectives are set and decisions<br />

made. This process not only acts as a catalyst by which various interest groups can imagine and<br />

formulate various futures, but when completed it serves as a contract setting out the<br />

responsibilities <strong>of</strong> the local users and the company/government (Berkes et al., 1991). Likewise,<br />

the practical management <strong>of</strong> forest resources should involve the active participation <strong>of</strong> local<br />

people, and their responsibility for the overall venture should be significant.<br />

The market values <strong>of</strong> NTFPs should be included in any financial analysis <strong>of</strong> forest management.<br />

Ethnobotanical literature on the region in which harvesting operations are to take place should<br />

be consulted. Traditional uses <strong>of</strong> species recorded in this literature can provide an idea <strong>of</strong> the


ange <strong>of</strong> products contained in the forest, and traditional management practices (such as<br />

management <strong>of</strong> fallows) can provide important "headstarts" in the design <strong>of</strong> appropriate<br />

silvicultural regimes.<br />

Harvesting operations should be staggered so that various areas are at various stages <strong>of</strong><br />

succession following disturbance and mature stands lie close to each other (WRI et al., 1992).<br />

By minimizing fragmentation, greater species diversity is maintained, and the likelihood <strong>of</strong> loss<br />

<strong>of</strong> NTFP species minimized.<br />

Refuges (composing 5-20% <strong>of</strong> the forest area), representing all forest types and particularly those<br />

with high species diversity and endemism, and corridors should be left between forest patches.<br />

Areas with unusual land fonns, geology or other physical features should be protected. Major<br />

saltlicks, feeding grounds, and other features which maintain particular wildlife populations<br />

should be preserved. Tree taxa that are important food sources for wild animals should be<br />

protected from harm during the logging operation and subsequent silvicultural regimes. Hunting<br />

by timber company employees or non-locals should be prohibited. Wildlife populations should<br />

be continuously monitored (Caldecott, 1988, Crome, 1991; John, A.D., 1992; ITIO Guidelines,<br />

1992).<br />

In addition to refuges established primarily to maintain wildlife and species diversity, areas<br />

containing significant numbers <strong>of</strong> non-timber forest product species should be avoided by logging<br />

operations (Lamb, 1991). Incentives and disincentives should be provided to loggers to insure<br />

that these measures are adequately carried out.<br />

Local communities should work with forest managers to design a schedule for logging operations<br />

that compliments their needs for employment, and incorporates the pre- and post-logging harvest<br />

<strong>of</strong> NTFPs. The planning <strong>of</strong> roads and skid trails should include provisions to not disrupt and<br />

perhaps facilitate the harvesting and transport <strong>of</strong> NTFPs.<br />

Silvicultural systems should not drastically alter the structural or species diversity <strong>of</strong> the forest.<br />

Enrichment plantings should be <strong>of</strong> native species, and should include multi-purpose species.<br />

The ecological sustainability <strong>of</strong> all NTFPs should be assessed, as far as is feasible, based on the<br />

plant part used, the floristic composition <strong>of</strong> the forest, the nature and intensity <strong>of</strong> harvesting, and<br />

the particular species or type <strong>of</strong> resource under exploitation (Peters, 1993), and should be<br />

monitored through a system <strong>of</strong> permanent sample plots.<br />

A diversity <strong>of</strong> non-timber forest products should be extracted from the forest to produce a system<br />

with seasonally complimentary harvests, and with reduced vulnerability to fluctuations in market<br />

demand for a particular product (Padoch et al., 1992; Reining and Heinzman, 1992; Phillips,<br />

1992).<br />

Where appropriate, local, regional and international NGOs should be enlisted to assist with the<br />

selection <strong>of</strong> species and development and expansion <strong>of</strong> external markets for a diversity <strong>of</strong><br />

products (Clay, 1992; Ziffer, 1992; Clay and Clement, 1993; Gentry, 1992). Where demand could<br />

exceed the capacity <strong>of</strong> existing harvesting systems to supply the market, alternative systems<br />

(including domestication <strong>of</strong> important species outside the forest area by local communities), must<br />

be employed (Cunningham, 1991).<br />

47


48<br />

The primary importance <strong>of</strong> non-timber forest products for subsistence and local markets should<br />

not be superseded by international demand for any NTFP product.<br />

Expropriation <strong>of</strong> resource use rights by powerful elites within these regions should be actively<br />

denied.<br />

The traditionally lopsided distribution <strong>of</strong> income from NTFPs - that is, the inverse relationship<br />

between the effort expended to produce a product and the income received - should be altered<br />

by the development or promotion <strong>of</strong> cooperatives, value-added processing and more equitable<br />

distribution <strong>of</strong> income along the chain <strong>of</strong> producers and marketers. The successful incorporation<br />

<strong>of</strong> NTFP extraction into forest management should not be dependent upon the physical and<br />

economic privation <strong>of</strong> extractors (Sizer, 1992; Browder, 1992).<br />

Traditional systems <strong>of</strong> resource management and regulation <strong>of</strong> harvesting should be employed,<br />

or their remnants built upon, to control resource extraction in the forest. Policing the removal <strong>of</strong><br />

NTFPs by local people is not a viable option. Collection <strong>of</strong> NTFPs by non-locals, however,<br />

should be discouraged, and methods employed devised by local communities.<br />

The long-tenn ownership and control <strong>of</strong> forest resources should clearly rest in the hands <strong>of</strong> local<br />

communities. In addition to the obvious ethical justification for land tenure, sustainable<br />

production is more likely to result from secure tenure in systems where the long-tenn, lowintensity<br />

harvesting <strong>of</strong>species from high-diversity forests can prove difficult and time-consuming.<br />

The nature and scale <strong>of</strong> timber harvesting operations should be adjusted to fit in with the existing<br />

NTFP harvest and trade patterns <strong>of</strong> local communities in cases where these are well-developed.<br />

Timber extraction should be seen as a compliment to these activities, rather than the primary<br />

objective <strong>of</strong> forest management.


Bibliography<br />

Abbiw, D.K. 1990. Useful Plants Of Ghana; West African Uses <strong>of</strong> Wild and Cultivated Plants.<br />

London: Intermediate Technology Publications.<br />

Abdillah, R. and C. Phillips. 1992. Preliminary Report on Rattan Damage Due to Selective Logging<br />

<strong>of</strong>Ridge Dipterocarp Forest in Sabah.<br />

Acquino, D.M. 1993. Non-Timber Forest Products Utilization in a Town in the Sierra Mountains.<br />

Paper presented at the International Association for the A Study <strong>of</strong> Common Property (IASCP)<br />

Fourth Annual Common Property Conference, The Philippines.<br />

Agyare, A.K. 1993. Partnership with indigenous peoples in sustainable natural resource use: a case<br />

in Ghana. paper presented at the Symposium for Indigenous People, The Pueblo <strong>of</strong> Zuni, New<br />

Mexico.<br />

Akerele, 0., V. Heywood and H. Synge (eds.). 1991. The Conservation <strong>of</strong> Medicinal Plants.<br />

Cambridge, U.K.: Cambridge <strong>University</strong> Press.<br />

Alcom, J. 1989. Process as resource: the traditional agricultural ideology <strong>of</strong> Bora and Huastec<br />

resource management in Amazonian before conquest; beyond ethnographic projection. In: D.<br />

Posey and W. Balee (OOs.). Resource Management in Amazonian: Indigenous and Folk<br />

Strategies. Advances in Economic Botany, 7.<br />

Alc<strong>of</strong>fi, J. 1990. Indigenous Agr<strong>of</strong>orestry Strategies Meeting fanners' Needs. In: A. Anderson (ed.)<br />

1990. Alternatives to Deforestation: Steps Toward Sustainable Use <strong>of</strong>the Amazon Rain Forest.<br />

New York: Columbia <strong>University</strong> Press.<br />

Alcorn, J. 1993. Indigenous peoples and conservation. Conservation Biology 7(2): 424-426.<br />

Alder, D. and T.J. Synnott. 1992. Permanent Sample Plot Techniques for Mixed Tropical Forest.<br />

Tropical Forestry Papers, No. 25. <strong>Oxford</strong> Forestry Institute, <strong>Oxford</strong>, U.K.<br />

Alencar, I.C. and N.P. Fernandes. 1978. Densenvolvimento de arvores nativas em ensaios de<br />

especies. Acta Amazonica 8(4): 523-541.<br />

Alencar, J.C. 1981. Estudos silviculturais de uma populacao naturel de Copai[era multijuga Hayne ­<br />

Legumunosae, na Amazonian Central, Acta Amazonica 11 (1): 3-11.<br />

Allegretti, M.H. 1990. Extractive Reserves: An Alternative for reconciling Development and<br />

Environmental Conservation in Amazonian. In: A. Anderson (ed.) 1990. Alternatives to<br />

Deforestation: Steps Toward Sustainable Use o/the Amazon Rain Forest. New York: Columbia<br />

<strong>University</strong> Press.<br />

Anderson, A. (ed.) 1990. Alternatives to Deforestation: Steps Toward Sustainable Use o/the Amazon<br />

Rain Forest. New York: Columbia <strong>University</strong> Press.<br />

49


50<br />

Anderson, A. 1990. Deforestation in Amazonian: Dynamics, Causes, and Alternatives. In: A.<br />

Anderson (ed.) 1990. Alternatives to Deforestation: Steps Toward Sustainable Use <strong>of</strong> the<br />

Amazon Rain Forest. New York: Columbia <strong>University</strong> Press.<br />

Anderson, A., P.H. May, and M.I. Balick. 1991. The Subsidy from Nature; Palm Forests, Peasantry,<br />

and Development on an Amazonian Frontier. New York: Columbia <strong>University</strong> Press.<br />

Amold, M. 1991. The Long Term Global Demand for and Supply <strong>of</strong> Wood. Forestry Commission<br />

Occasional Paper, 36.<br />

Balee, W. 1988. Indigenous adaptation to Amazonian palm forests. Principes 32(2): 47-54.<br />

Balee, W. and A, Gely. 1989. Managed forest succession in Amazonian: the Ka'apor case. In: D.<br />

Posey and W. Balee (eds.). Resource Management in Amazonian: Indigenous and Folk<br />

Strategies. Advances in Economic Botany, 7.<br />

Balee, W. 1989. The culture <strong>of</strong> Amazonian forests. In: D. Posey and W. Balee (eds.). Resource<br />

Management in Amazonian: Indigenous and Folk Strategies. Advances in Economic Botany, 7.<br />

Balick, M.J. and R. Mendelssohn. 1992. Assessing the economic value <strong>of</strong> traditional medicines from<br />

tropical rain forests. Conservation Biology 6(1): 128-130.<br />

Barbier, E.B. 1987. The concept <strong>of</strong> sustainable economic development. Environmental Conservation<br />

14: 101-110.<br />

Bazzaz, F.A. 1991. Regeneration <strong>of</strong>tropical forests: physiological responses <strong>of</strong>pioneer and secondary<br />

species. in: A. Gomez-Pompa, T.C. Whitmore and M. Hadley (eds.) Rain Forest Regeneration<br />

and Management. Paris: UNESCO and The Parthenon Publishing Group Limited.<br />

Bennett, B. 1992. Plants and people <strong>of</strong> the Amazonian rainforests: the role <strong>of</strong> ethnobotany in<br />

sustainable development. Bioscience 42(8).<br />

Berkes, F., P. George, and R.J. Preston. 1991. Co-management: the evolution in theory and practice<br />

<strong>of</strong> the joint administration <strong>of</strong> living resources. Alternatives 18(2):12-39<br />

Blockhus, I.M., M. Dillenbeck, I.A. Sayer, and P. Wegge. 1992. Conserving Biological Diversity in<br />

Managed Tropical Forests. The IUCN Forest Conservation Programme, Cambridge, U.K.<br />

Boom, B. 1989. Use <strong>of</strong> plant resources by the Chacabo. In: D. Posey and W. Balee (eds.). Resource<br />

Management in Amazonian: Indigenous and Folk Strategies. Advances in Economic Botany, 7<br />

Browder, J.O. 1986. Logging The Rainforest: A Political Economy <strong>of</strong>Timber Extraction and Unequal<br />

Exchange in the Brazilian Amazon. <strong>University</strong> <strong>of</strong> Pennsylvania Dissertation.<br />

Browder, J.O. 1992. The limits <strong>of</strong> extractivism; tropical forest strategies beyond extractive reserves.<br />

Bioscience 42(3): 174-182.


Brown, K.S. and G.G. Brown. 1992. Habitat alteration and species loss in Brazilian forests. In: T.e.<br />

Whitmore and J.A. Sayer (eds.) Tropical Deforestation and Species Extinction. London:<br />

Chapman and HalL<br />

Brown, N. and M. Press. 1992. Logging rainforest the natural way? New Scientist, 14 March.<br />

Brown, N.D. and T.C. Whitmore. 1992. Do dipterocarp seedlings really partition tropical rain forest<br />

gaps? Philosophical Transactions <strong>of</strong> the Royal Society 335: 369-378.<br />

Burch, W.B. 1983. Summary: Toward Social Forestry in the Tropics. In: F. Mergen (ed.) Tropical<br />

Forests: Utilization and Conservation. Proceedings <strong>of</strong> an International Symposium held at Yale<br />

<strong>University</strong>, New Haven, CT.<br />

Bums, D. 1986. Runway and Treadmill Deforestation: Reflections on the Economics <strong>of</strong> Forest<br />

Development in the Tropics. London: lIED and Earthscan Publications ltd.<br />

Caldecott, J. 1988. Hunting and Wildlife Management in Sarawak. The mCN Tropical Forest<br />

Programme, Cambridge, UK.<br />

Chin, L.T. 1973. Occurrence <strong>of</strong> seeds in virgin forest top soil with particular reference to secondary<br />

species in Sabah. The Malaysian Forester 36(3):185-193.<br />

Clay, J. 1992. Some general principles and strategies for developing markets in North America and<br />

Europe for nontimber forest products. In: M. Plotkin and L. Famolare (eds.). Sustainable<br />

Harvest and Marketing <strong>of</strong>Rain Forest Products. Washington, D.C.: Island Press.<br />

Clay, J. 1992. Some general principles and strategies for developing markets in North America and<br />

Europe for non-timber forest products: lessons from Cultural Survival Enterprises, 1989-1990.<br />

In: D.C. Nepstad and S. Schwartzrnan (eds.). Non-Timber Products from Tropical Forests;<br />

Evaluation <strong>of</strong> a Conservation and Development Strategy. Advances in Economic Botany,<br />

Volume 9. The New York Botanical Garden, New York.<br />

Clay, J. and C.R. Clement. 1993. Selected Species and Strategies to Enhance Income Generation<br />

from Amazonian Forests. FAO Forestry Paper (Draft), Rome, Italy.<br />

Cleary, D.M. 1992. Overcoming socio-economic and political constraints to "wise forest<br />

management": lessons from the Amazon. In: F. Miller and K.L. Adam (eds.). 1992. Wise<br />

Management o/Tropical Forests. Proceedings <strong>of</strong> the <strong>Oxford</strong> Forestry Institute Conference on<br />

Tropical Forests.<br />

Crorne, F.H.J. 1991. Wildlife conservation and rain forest management - examples from North East<br />

Queensland. In: A. Gomez-Pompa, T.C. Whitmore and M. Hadley (eds.) Rain Forest<br />

Regeneration and Management. Paris: UNESCO and The Parthenon Publishing Group Limited.<br />

Cultural Survival. 1992. Cupuacu (Theobroma grandijlorum). Product Descriptions. Cambridge, MA.<br />

51


52<br />

Cunningham, A.B. 1989. Indigenous plant use: balancing human needs and resources. In: B. Huntley<br />

(eel.) Conserving Biological Diversity in Southern Africa. <strong>Oxford</strong>: <strong>Oxford</strong> <strong>University</strong> Press.<br />

Cunningham, A.B. 1990. African Medicinal Plants: Setting Priorities at the Interface Between<br />

Conservation and Primary Healthcare.Report for WWF Project 3331.<br />

Cunningham, A.B. 1991. Development <strong>of</strong> a conservation policy on commercially exploited medicinal<br />

plants: a case study from southern Africa. In: O. Akerele, V. Heywood and H. Synge (eds.). The<br />

Conservation <strong>of</strong> Medicinal Plants. Cambridge, U.K.: Cambridge <strong>University</strong> Press.<br />

Cunningham, A.B. 1992. People, Park and Plant Use: Research and Recommendations for Multiple­<br />

Use Zones and Development Alternatives Around Bwindi-Impenetrable National Park, Uganda.<br />

Report prepared for CARE-International, Kampala, Uganda.<br />

Cunningham, A.B. and F.T. Mbenkum. 1993. Medicinal Bark in International Trade: a Case Study<br />

<strong>of</strong> the Afromontane Tree Prunus Africana. Report to WWF International.<br />

Dahaban, Z., M. Nordin and E.L. Bennett. 1992. Immediate effects on wildlife is selective logging<br />

in a hill dipterocarp forest in Sarawak: mammals. Department <strong>of</strong> Zoology, <strong>University</strong><br />

Kebangsaan, Malaysia and Wildlife Conservation Society.<br />

Davies, A.G. 1987. The Gola Forest Reserves, Sierra Leone; Wildlife Conservation and Forest<br />

Management. IUCN Tropical Forest Programme, Cambridge, U.K.<br />

Davies, A.G. and P. Richards. 1991. Rain Forest in Mende Life; resource and subsistence strategies<br />

in rural communities around the Gola North forest reserve. A report to ESCOR, UK Overseas<br />

Development Administration.<br />

Dawkins, H.C. 1958. The Management <strong>of</strong> Natural Tropical High-Forest with Special Reference to<br />

Uganda. Imperial Forestry Institute Paper, No. 34. <strong>University</strong> <strong>of</strong> <strong>Oxford</strong>, <strong>Oxford</strong>.<br />

De Beer, J.H. and M.J. McDennott. 1989. The Economic Value <strong>of</strong>Non-Timber Forest Products in<br />

Southeast Asia with Emphasis on Indonesia, Malaysia and Thailand. Netherlands Committee for<br />

IUCN, Amsterdam.<br />

De Graaf, N.R. 1986. A Silvicultural System for Natural Regeneration <strong>of</strong> Tropical Rain Forest in<br />

Suriname. The Agricultural <strong>University</strong>, Wageningen, The Netherlands.<br />

De Graaf and R.L.H. Pools. 1990. The Celos Management System: A Polycyclic Method for<br />

Sustained Timber Production in South American Rainforest. In: A. Anderson (ed.) 1990.<br />

Alternatives to Deforestation: Steps Toward Sustainable Use <strong>of</strong>the Amazon Rain Forest. New<br />

York: Columbia <strong>University</strong> Press.<br />

De Graaf, N.R. 1991. Managing natural regeneration for sustained timber production in Suriname:<br />

the Celos Silvicultural and Harvesting System. In: A. Gomez-Pompa, T.C. Whitmore and M.<br />

Hadley (eds.) Rain Forest Regeneration and Management. Paris: UNESCO and The Parthenon<br />

Publishing Group Limited.


De Jong, W.D. and R. Mendelssohn. 1992. Managing the Non-Timber Forest Products <strong>of</strong>Southeast<br />

Asia. Paper prepared for the World Bank, Southeast Asia Division.<br />

Denslow, J.S. 1980. Gap partitioning among tropical rainforest trees. Tropical Succession: 47-55.<br />

Dixon, A., H. Roditi, and L. Silvennan. 1991. From Forest to Market: A Feasibility Study <strong>of</strong> the<br />

Development <strong>of</strong>Selected Non-Timber Forest Products from Borneo for the U.S. Market. Project<br />

Borneo, Cambridge, MA.<br />

Donovan, R. 1993. Independent Forestry Certification as a Tool for Improving Forestry Practices.<br />

Paper presented at the First Euro-Seminar on Promoting a Trade in Sustainably Produced<br />

Timber, EC, Brussels.<br />

Dwyer, I.D. 1951. The Central American, West Indian, and South American Species <strong>of</strong> Copaijera<br />

(Caesalpiniaceae). Brittonia 7(3): 143-172.<br />

Dykstra, D.P. and R. Heinrich. 1992. Sustaining tropical forests through environmentally sound<br />

harvesting practices. Unasylva 43 (169): 9-15.<br />

Elisabetsky, E. 1991. Sociopolitical, economic and ethical issues in medicinal plant research. Journal<br />

<strong>of</strong> Ethnophannacology 32.<br />

Estes, J.A., D.O. Duggins and G.B. Rathbun. 1989. The ecology <strong>of</strong> extinctions in jelp forest<br />

communities. Conservation Biology 3(3): 252-264.<br />

Ewel, J. 1980. Tropical succession: manifold routes to maturity. Tropical Succession: 2-7.<br />

Ewel, J. 1983. Environmental Implications <strong>of</strong> Utilization. In: F. Mergen (ed.) Tropical Forests:<br />

Utilization and Conservation. Proceedings <strong>of</strong> an International Symposium held at Yale<br />

<strong>University</strong>, New Haven, CT.<br />

Falconer, J. 1990. The Major Significance oj'Minor' Forest Products; The Local Use and Value <strong>of</strong><br />

Forests in the West African Humid Forest Zone. FAO Community Forestry Note No. 6, Rome.<br />

Falconer, J. 1992. A study <strong>of</strong> the non-timber forest products <strong>of</strong> Ghana's forest zone. In: The<br />

Rainforest Harvest; Sustainable Strategiesfor Saving the Tropical Forests? Friends <strong>of</strong>the Earth,<br />

London.<br />

Falconer, J. 1992. Non-Timber Forest Products in Southern Ghana. ODA Forestry Series, No. 2,<br />

London.<br />

Fanshawe, D.B. 1947. Studies <strong>of</strong> the Trees <strong>of</strong> British Guiana; Crabwood. Tropical Woods 90: 30-40.<br />

FAO. 1991. Non-wood Forest Products: The Way Ahead. FAO Forestry Paper 97, Rome.<br />

53


54<br />

Fearnside, P. 1990. Predominant land uses in Brazilian Amazon. In: A. Anderson (ed.) 1990.<br />

Alternatives to Deforestation: Steps Toward Sustainable Use <strong>of</strong>the Amazon Rain Forest. New<br />

York: Columbia <strong>University</strong> Press.<br />

Forest Stewardship Council. 1992. Draft Principles and Criteria for Forest Management.<br />

Fox, J.E.D. 1968. Logging damage and the influence <strong>of</strong> climber cutting prior to logging in the<br />

lowland Dipterocarp forest <strong>of</strong> Sabah. The Malaysian Forester 31(4): 326-347.<br />

Friends <strong>of</strong> the Earth. 1992. The Rainforest Harvest: Sustainable Strategies for Saving the Tropical<br />

Forests? London.<br />

Fujisaka, S. and E. Wollenberg. 1991. From forest to agr<strong>of</strong>orest and logger to agr<strong>of</strong>orester: a case<br />

study. Agr<strong>of</strong>orestry Systems 14: 113-129.<br />

Gane, M. 1992. Sustainable forestry. Commonwealth Forestry Review 71(2): 83-90.<br />

Gentry, A.H. and R. Vasquez. 1988. Where have all the Ceibas Gone? A Case History <strong>of</strong><br />

Mismanagement <strong>of</strong> a Tropical Forest Resource. Forest Ecology and Management 23: 73-76.<br />

Gentry, A.H. 1992. New nontimber forest products from western South America. In: M. Plotkin and<br />

L. Famolare (eds.). Sustainable Harvest and Marketing <strong>of</strong>Rain Forest Products. Washington,<br />

D.C.: Island Press.<br />

Godoy, R. and R. Lubowski. 1992. Guidelines for the Economic Valuation <strong>of</strong> Nontimber tropical­<br />

Forest Products. Current Anthropology 33(4): 423-430.<br />

GOOoy, R., N.L.V. Brokaw, and D. Wilkie. 1992. The Effects <strong>of</strong> Economic Development on the<br />

Extraction <strong>of</strong> Non-Timber Tropical Forest Products: Hypotheses, Methods, and Information<br />

Requirements. Harvard Institute for International Development, Cambridge, MA.<br />

Gomez-Pompa, A., C. Vasquez-Yanes, and S. Guevara. 1972. The tropical forest: a nonrenewable<br />

resource. Science 177: 762-765.<br />

Gomez-Pompa, A. and A. Kaus. 1990. Traditional Management <strong>of</strong> Tropical Forests in Mexico. In:<br />

A. Anderson (ed.) 1990. Alternatives to Deforestation: Steps Toward Sustainable Use <strong>of</strong> the<br />

Amazon Rain Forest. New York: Columbia <strong>University</strong> Press.<br />

Gomez-Pompa, A. 1991. Learning from traditional ecological knowledge: insights from Mayan<br />

silviculture. in: A. Gomez-Pompa, T.C. Whitmore and M. Hadley (eds.) Rain Forest<br />

Regeneration and Management. Paris: UNESCO and The Parthenon Publishing Group Limited.<br />

Gomez-Pompa, A. and F.W. Burley. 1991. The management <strong>of</strong> natural tropical forests. In: A.<br />

Gomez-Pompa, T.e. Whitmore and M. Hadley (eds.) Rain Forest Regeneration and<br />

Management. Paris: UNESCO and The Parthenon Publishing Group Limited.


Gomez-Pompa, A., T.C. Whitmore and M. Hadley (eds.). 1991. Rain Forest Regeneration and<br />

Management. Paris: UNESCO and The Parthenon Publishing Group Limited.<br />

Grajal, Alejandro, Wildlife Conservation Society, personal communication, 1993.<br />

Grimes, A., S. Loomis, P. Jahnige, M. Bumham, K. Onthank, R. Alarcon, W. Palacios, C. Ceron, D.<br />

Neill, M. Balick, B, Bennett, and R. Mendelssohn. 1993. Valuing the Rain Forest: The<br />

Economic Value <strong>of</strong> Non-Timber Forest Products in Ecuador. Yale School <strong>of</strong> Forestry and<br />

Environmental Studies and The New York Botanical Garden.<br />

Guenther, E. 1948. Essential Oils. Van Nostrad Co. Publishers.<br />

Hansen, A.J., T.A. Spies, F.J. Swanson and J.L. Ohmann. 1991. Conserving Biodiversity in Managed<br />

Forests; Lessons from natural forests. Bioscience 41(6)382-392.<br />

Hart, R.D. 1980. A natural ecosystem analog approach to the design <strong>of</strong> a successional crop system<br />

for tropical forest environments. Tropical Succession: 73-82.<br />

Hartshorn, G.S. 1980. Neotropical Forest Dynamics. Tropical Succession: 23-30.<br />

Hartshorn, G. 1990. Natural Forest Management by the Yanesha Forestry Cooperative. In: A.<br />

Anderson (ed.) 1990. Alternatives to Deforestation: Steps Toward Sustainable Use <strong>of</strong> the<br />

Amazon Rain Forest. New York: Columbia <strong>University</strong> Press.<br />

Hecht, S. and Cockbum, A. 1989. The Fate <strong>of</strong>the Forest; Developers, Destroyers and Defenders <strong>of</strong><br />

the Amazon. London: Verso Press.<br />

Hendrison, J. 1990. Damage-Controlled Logging in Managed Rain Forest in Suriname. The<br />

Wageningen Agricultural <strong>University</strong>, The Netherlands.<br />

Heywood, V.H. and S.N. Stewart. 1992. Species extinction in tropical forests. In: T.C. Whitmore and<br />

J.A. Sayer (eds.) Tropical Deforestation and Species Extinction. London: Chapman and Hall.<br />

Hidalgo, R.C. 1992. The Tagua Initiative in Ecuador: a community approach to tropical rain forest<br />

conservation and development. In: M. Plotkin and L. Famolare (eds.). Sustainable Harvest and<br />

Marketing <strong>of</strong>Rain Forest Products. Washington, D.C.: Island Press.<br />

Husain, A. 1991. Economic aspects <strong>of</strong> exploitation <strong>of</strong> medicinal plants. In: O. Akerele, V. Heywood<br />

and H. Synge (eds.). The Conservation <strong>of</strong> Medicinal Plants. Cambridge, U.K.: Cambridge<br />

<strong>University</strong> Press.<br />

Irvine, D. 1989. Succession management and resource distribution in an Amazonian rain forest. In:<br />

D. Posey and W. Balee (eds.). Resource Management in Amazonian: Indigenous and Folk<br />

Strategies. Advances in Economic Botany, 7.<br />

ITIO. 1990. Guidelines for the Sustainable Management <strong>of</strong>Natural Tropical Forests.<br />

55


56<br />

ITIO. 1992. Draft Guidelines for Conserving Biological Diversity in Forests Managedfor Timber.<br />

IUCN Inter-Commission Task Force on Indigenous Peoples. 1993. Indigenous Peoples and Strategies<br />

for Sustainability. Workshop report, Gland, Switzerland.<br />

Jain, S.K. 1992. Folk Beliefs in Conservation <strong>of</strong> Nature and Culture. Paper presented at the<br />

International Congress <strong>of</strong> Ethnobiology, Mexico.<br />

John, A.D. 1988. Effects <strong>of</strong> "selective" timber extraction on rainforest structure and composition and<br />

some consequences for frugivorous and folivores. Biotropica 20(1): 31-37.<br />

John, A.D. 1992. Species conservation in managed tropical forests. In: T.C. Whitmore and J.A. Sayer<br />

(OOs.) Tropical Deforestation and Species Extinction. London: Chapman and Hall.<br />

John, A.D. 1991. Responses <strong>of</strong> Amazonian rain forest birds to habitat modification. Journal <strong>of</strong><br />

Tropical Ecology 7: 417-437.<br />

John, R.J. 1992. The influence <strong>of</strong> deforestation and selective logging operations on plant diversity<br />

in Papua New Guinea. In: T.C. Whitmore and J.A. Sayer (eds.) Tropical Deforestation and<br />

Species Extinction. London: Chapman and Hall.<br />

Jonson, T. and P. Lindgren. 1990. Logging Technology for Tropical Forests - for or against? Report<br />

from the ITTO pre-project "Improvement <strong>of</strong> Harvesting Systems for the Sustainable<br />

Management <strong>of</strong>Tropical Rain Forests". The Forest Operations Institute 'Skogsarbeten', Sweden.<br />

Katz-Miller, S. 1993. High hope hanging on a "useless" vine. New Scientist, 16 January.<br />

Kemp, R.H., G. Namkoong and F.R. Wadsworth. 1993. Conservation <strong>of</strong> Genetic Resources in<br />

Tropical Forest Management; Principles and Concepts. FAO Forestry Paper 107. FAO, Rome,<br />

Italy.<br />

Kraner, R., R. Realy and R. Mendelssohn. 1992. Forest Valuation. In: N.P. Shanna (ed.) Managing<br />

the World' s Forests. Kendall/Hunt.<br />

Lamb, D. 1991. Combining traditional and commercial uses <strong>of</strong> rain forests. Nature and Resources<br />

27(2):3-11.<br />

Lawton, R.M. 1989. The exploitation <strong>of</strong> the African rain forest and man's impact. In: H. Leith and<br />

M.J.A. Werger (OOs.). Tropical Rain Forest Ecosystems. Amsterdam: Elsevier Science<br />

Publishers.<br />

Leslie, A.J. 1987. The economic feasibility <strong>of</strong> natural management <strong>of</strong> tropical forests. In: F. Mergen<br />

and J.R. Vincent (OOs.). Natural Management <strong>of</strong> Tropical Moist Forests; Silvicultural and<br />

Management Prospects <strong>of</strong> Sustained Utilization. Yale <strong>University</strong>, School <strong>of</strong> Forestry and<br />

Environmental Studies, New Haven, CT.


Lock, I.M. 1986. Plant-animal interactions. In: G.W. Lawson (00.) Plant Ecology in West Africa.<br />

New York: John Wiley & Sons Ltd.<br />

Malhorta, K.C., M. P<strong>of</strong>fenberger, A. Bhattacharya and D. Dev. 1991. Rapid appraisal methodology<br />

trials in Southwest Bengal: assessing natural forest regeneration patterns and non-wood forest<br />

product harvesting practices. Forests, Trees and People Newsletter No. 15/16.<br />

Martini, A.M.Z., N. de A. Rosa, and C. UhI. 1992. A First Attempt to Predict Amazonian Tree<br />

Species Threatened by Logging Pressure. Instituto do Homen e Meio Ambiente da Amazonian<br />

(IMAZON), Belem, Brazil.<br />

Martins, P. Canadian International Development Agency, pers. comm., 1992.<br />

Mather, A.S. 1990. Global Forest Resources. Portland, Oregon: Timber Press.<br />

McNeely, I.A. 1988. Economics and Biological Diversity; Developing and Using Economic<br />

Incentives to Conserve Biological Resources. mCN, Gland, Switzerland.<br />

Meijer, W. 1970. Regeneration <strong>of</strong> tropical lowland forest in Sabah, Malaysia forty years after<br />

logging. The Malaysian Forester 33(3): 204-228.<br />

Mergen, F. (ed.). 1983. Tropical Forests: Utilization and Conservation. Proceedings <strong>of</strong> an<br />

International Symposium held at Yale <strong>University</strong>, New Haven, CT.<br />

Mergen, F. and J.R. Vincent (eds.). 1987. Natural Management <strong>of</strong> Tropical Moist Forests;<br />

Silvicultural and Management Prospects <strong>of</strong> Sustained Utilization. Yale <strong>University</strong>, School <strong>of</strong><br />

Forestry and Environmental Studies, New Haven, cr.<br />

Miller, F. and K.L. Adam (eds.). 1992. Wise Management <strong>of</strong> Tropical Forests. Proceedings <strong>of</strong> the<br />

<strong>Oxford</strong> Forestry Institute Conference on Tropical Forests.<br />

Milliken, W., R.P. Miller, S.R. Pollard, and E.V. Wandelli. 1992. Ethnobotany o/the Waimiri Atroari<br />

Indians <strong>of</strong>Brazil. Royal Botanic Gardens, Kew.<br />

Mori, S. 1992. The Brazil nut industry - past, present, and future. In: M. Plotkin and L. Famolare<br />

(eds.). Sustainable Harvest and Marketing <strong>of</strong>Rain Forest Products. Washington, D.C.: Island<br />

Press.<br />

Nair, C.T.S. 1991. A comparative account <strong>of</strong> silviculture in the tropical wet evergreen forests <strong>of</strong><br />

Kerala, Andaman Islands and Assam. In: A. Gomez-Pompa, T.C. Whitmore and M. Hadley<br />

(eds.) Rain Forest Regeneration and Management. Paris: UNESCO and The Parthenon<br />

Publishing Group Limited.<br />

National Research Council. 1992. Conserving Biodiversity; A Research Agenda for Development<br />

Agencies. Washington, D.C.: National Academy Press.<br />

57


58<br />

Neiring, W.A. 1987. Vegetation dynamics (succession and climax) in relation to plant community<br />

management. Conservation Biology 1(4): 287-295.<br />

Nepstad, D.C., C. Uhl and E.A.S. Serrao. 1991. Recuperation <strong>of</strong> a degraded Amazonian landscape:<br />

forest recovery and agricultural restoration. Ambio 20(6): 248-255.<br />

Nepstad, D.C. and I.F. Brown, V. Viana and L. Luz. 1992. Biotic impoverishment <strong>of</strong> Amazonian<br />

forests by rubber tappers, loggers and cattle ranchers. In: D.C. Nepstad and S. Schwartzman<br />

(eds.). Non-Timber Products from Tropical Forests; Evaluation <strong>of</strong> a Conservation and<br />

Development Strategy. Advances in Economic Botany, Volume 9. The New York Botanical<br />

Garden, New York.<br />

Nepstad, D.C. and S. Schwartzman (eds.). 1992. Non-Timber Products from Tropical Forests;<br />

Evaluation <strong>of</strong> a Conservation and Development Strategy. Advances in Economic Botany,<br />

Volume 9. The New York Botanical Garden, New York.<br />

North Queensland Department <strong>of</strong> Forestry. Guidelines for the Selective Logging <strong>of</strong>Rainforest Areas<br />

in North Queensland State Forests and Timber Reserves.<br />

Ocana-Vidal, J. 1992. Natural forest management with strip clear-cutting. Unasylva 43(169): 24-27.<br />

Oldfield, M. 1989. The Value <strong>of</strong>Conserving Genetic Resources. Massachusetts: Sinauer Associates,<br />

Inc. Publishers.<br />

Othole, A. and R. Anyon. 1993. A Tribal Perspective on Traditional Cultural Property Consultation.<br />

Cultural Preservation Program. The Pueblo <strong>of</strong> Zuni, New Mexico.<br />

Padoch, C., T.C. Jessup, H. Soedjito and K. Kartawinata. 1991. Complexity and conservation <strong>of</strong><br />

medicinal plants: anthropological cases from Peru and Indonesia. In: O. Akerele, V. Heywood<br />

and H. Synge (eds.). The Conservation <strong>of</strong> Medicinal Plants. Cambridge, U.K.: Cambridge<br />

<strong>University</strong> Press.<br />

Padoch, C. and De Jong, W. 1992. Diversity, variation, and change in Ribereno agriculture. In: K.H.<br />

Redford and C. Padoch (eds.). Conservation <strong>of</strong>Neotropical Forests; Working from Traditional<br />

Resource Use. New York: Columbia <strong>University</strong> Press.<br />

Padoch, C. 1992. Marketing <strong>of</strong> non-timber forest products in western Amazonian: general<br />

observations and research priorities. In: D.C. Nepstad and S. Schwartzman (eds.). Non-Timber<br />

Products from Tropical Forests; Evaluation <strong>of</strong> a Conservation and Development Strategy.<br />

Advances in Economic Botany, Volume 9. The New York Botanical Garden, New York.<br />

Parren, M. 1992. Non-timber forest products: weighing the benefits against the costs. In: J. Hummel<br />

and M. Parren (oos.). Forests a Growing Concern. Proceedings <strong>of</strong> the XIXth International<br />

Forestry Students Symposium, Wageningen, The Netherlands.<br />

Pearce, D. 1991. Assessing the Returns to the Economy and to Society from Investments in Forestry.<br />

Forestry Commission Occasional Paper, 47.


Pearce, D. and S. Puroshothaman. 1993. Protecting Biological Diversity: The Economic Value <strong>of</strong><br />

Pharmaceutical Plants. Prepared for T. Swanson (ed.) Biodiversity and Botany: The Values <strong>of</strong><br />

Medicinal Plants. Forthcoming, CSERGE.<br />

Peluso, N.L. 1992. Rich Forests, Poor People; Resource Control and Resistance in Java. Berkeley:<br />

<strong>University</strong> <strong>of</strong> California Press.<br />

Pendelton, L.H. 1992. Trouble in paradise: practical obstacles to nontimber forestry in Latin America.<br />

In: M. Plotkin and L. Famolare (eds.). Sustainable Harvest and Marketing <strong>of</strong> Rain Forest<br />

Products. Washington, D.C.: Island Press.<br />

Peters, C.M., A.H. Gentry and R.O. Mendelssohn. 1989. Valuation <strong>of</strong> an Amazonian rainforest.<br />

Nature 339: 655-656.<br />

Peters, C.M., M.J. Balick, and F. Kahn. 1989. Oligarchic forests <strong>of</strong> economic plants in Amazonian:<br />

utilization and conservation <strong>of</strong> an important tropical resource. Conservation Biology 3(4): 341­<br />

349.<br />

Peters, C.M. 1993. Ecology and Exploitation <strong>of</strong>Non-timber Tropical Forest Resources: A Primer on<br />

Sustainability. Draft report, Biodiversity Support Program, Washington, D.C.<br />

Phillips, O. 1993. The potential for harvesting fruits in tropical rainforests: new data from Amazonian<br />

Peru. Biodiversity and Conservation 2: 18-38.<br />

Phillips, O. and Gentry, A.H. 1993. The useful plants <strong>of</strong> Tambopata, Peru: I. statistical hypotheses<br />

tests with a new quantitative technique. Economic Botany 47(1): 15-32.<br />

Plotkin, M. and L. Famolare. 1992. Sustainable Harvest and Marketing <strong>of</strong> Rain Forest Products.<br />

Washington, D.C.: Island Press.<br />

Poole, P. 1993. Land-Based Communities and Biodiversity Conservation: A Survey <strong>of</strong> Current<br />

Activities in Relation to Implementation <strong>of</strong> the Biodiversity Convention. Prepared for the<br />

Biodiversity Office, Environment, Canada.<br />

Poore, D., P. Burgess, J. Palmer, S. Rietbergen and T. Synnott. 1989. No Timber Without Trees;<br />

Sustainability in the Tropical Forest. London: earthscan Publications Ltd.<br />

Poore, D. and J. Sayer. 1991. The Management <strong>of</strong> Tropical Moist Forest Lands: Ecological<br />

Guidelines. The IUCN Forest Conservation Programme, Cambridge, U.K.<br />

Posey, D.A. 1990. The Application <strong>of</strong> Ethnobiology in the Conservation 0/ Dwindling Natural<br />

Resources: Lost Knowledge or Options/or the Survival <strong>of</strong>the Planet. Proceedings from the First<br />

International Congress <strong>of</strong> Ethnobiology, Brazil: 47-61.<br />

Posey, D.A. 1992. Traditional knowledge, conservation and "the rain forest harvest". In: M. Plotkin<br />

and L. Famolare (eds.). Sustainable Harvest and Marketing <strong>of</strong> Rain Forest Products.<br />

Washington, D.C.: Island Press.<br />

59


60<br />

Posey, D.A. and W. Balee (eds.). 1989. Resource Management in Amazonian: Indigenous and Folk<br />

Strategies. Advances in Economic Botany, 7.<br />

Prescott-Allen, R. and C. 1982. What's Wildlife Worth? Economic Contributions o/Wild Plants and<br />

Animals to Developing Countries. London: Earthscan Publications Ltd.<br />

Principe, P.P. 1991. Valuing the biodiversity <strong>of</strong> medicinal plants. In: 0. Akerele, V. Heywood and<br />

H. Synge (eds.). The Conservation <strong>of</strong>Medicinal Plants. Cambridge, U.K.: Cambridge <strong>University</strong><br />

Press.<br />

Putz, F.E. 1984. How trees avoid and shed lianas. Biotropica 16(1): 19-23.<br />

Putz, F.E. and N.V.L. Brokaw. 1989. Sprouting <strong>of</strong> broken trees on Barro Colorado Island, Panama.<br />

Ecology 70(2): 508-512.<br />

Record, S.J. and C.D. Mel!. 1924. Timbers o/Tropical America. New Haven, CT: Yale <strong>University</strong><br />

Press.<br />

Redford, K.H. 1991. Hunting in Neotropical Forests: A Subsidy from Nature. Presented at "Nutrition<br />

in Tropical Forests", UNESCO, Paris.<br />

Redford, K.H. and C. Padoch. 1992. Conservation o[Neotropical Forests; Working from Traditional<br />

Resource Use. New York: Columbia <strong>University</strong> Press.<br />

Redford, K.H. and A.M. Stearman. 1993. Forest-dwelling native Amazonians and the conservation<br />

<strong>of</strong> biodiversity: interests n common or in collision? Conservation Biology 7(2): 248-255.<br />

Reid, W.V. 1992. How many species will there be? In: T.C. Whitmore and J.A. Sayer (eds.) Tropical<br />

Deforestation and Species Extinction. London: Chapman and Hall.<br />

Reining, C. and R. Heinzman. 1992. Nontimber forest products in the Peten, Guatemala: why<br />

extractive reserves are critical for both conservation and development. In: M. Plotkin and L.<br />

Famolare (eds.). Sustainable Harvest and Marketing <strong>of</strong>Rain Forest Products. Washington, D.C.:<br />

Island Press.<br />

Reining, C., R.M. Heinzman, M.C. Madrid, S. Lopez, and A. Solorzano. 1992. Non-Timber Forest<br />

Products <strong>of</strong> the Maya Biosphere reserve, Peten, Guatemala. Conservation International,<br />

Washington, D.C.<br />

Repetto, R. and M. Gillis. 1988. Public Policies and the Misuse <strong>of</strong> Forest Resources. Cambridge:<br />

Cambridge <strong>University</strong> Press.<br />

Richards, M. 1993. The potential <strong>of</strong> non-timber forest products in sustainable natural forest<br />

management in Amazonian. Commonwealth Forestry Review 72(1): 21-27.<br />

Rosenberg, D.B. and S.M. Freedman. 1984. Application <strong>of</strong> a model <strong>of</strong> ecological succession to<br />

conservation and land-use management. Environmental Conservation 11(4): 323-330.


Salick, J. 1992. Amuesha forest use and management: an integration <strong>of</strong> indigenous use and natural<br />

forest management. In: K.H. Redford and C. Padoch (OOs.). Conservation <strong>of</strong> Neotropical<br />

Forests; Working from Traditional Resource Use. New York: Columbia <strong>University</strong> Press.<br />

Salick, J. 1992. The sustainable management <strong>of</strong> nontimber rain forest products in the Si-a-Paz Peace<br />

Park, Nicaragua. In: M. Plotkin and L. Famolare (OOs.). Sustainable Harvest and Marketing <strong>of</strong><br />

Rain Forest Products. Washington, D.C.: Island Press.<br />

Saulei, S. 1984. Natural regeneration following clear-fell logging operations in the Gogol Valley,<br />

Papua New Guinea. Ambio 13(5-6): 351-354.<br />

Sayer, J. 1991. Rainforest Buffer Zones; Guidelines for Protected Area Managers. IUCN - The World<br />

Conservation Union, Forest Conservation Programme, Cambridge, U.K.<br />

Sayer, J.A. and Whitmore, T.C. 1991. Tropical moist forests: destruction and species extinction.<br />

Biological Conservation 55: 199-213.<br />

Schmidt, R.C. 1991. Tropical rain forest management: a status report. In: A. Gomez-Pompa, T.C.<br />

Whitmore and M. Hadley (eds.) Rain Forest Regeneration and Management. Paris: UNESCO<br />

and The Parthenon Publishing Group Limited.<br />

Schultes, R.E. 1992. Ethnobotany and technology in the northwest Amazon: a partnership. In: M.<br />

Plotkin and L. Famolare (eds.). Sustainable Harvest and Marketing <strong>of</strong> Rain Forest Products.<br />

Washington, D.C.: Island Press.<br />

Schwartzman, S. 1992. Land distribution and the social costs <strong>of</strong> frontier development in Brazil: social<br />

and historical context <strong>of</strong> extractive reserves. In: D.C. Nepstad and S. Schwartzman (OOs.). Non­<br />

Timber Products from Tropical Forests; Evaluation <strong>of</strong> a Conservation and Development<br />

Strategy. Advances in Economic Botany, Volume 9. The New York Botanical Garden, New<br />

York.<br />

Sheldon, J.W., M.J. Balick and S.A. Laird. 1993. Medicinal Plants: Can Utilization and Conservation<br />

Co-Exist? Draft, Advances in Economic Botany. The New York Botanical Garden, New York.<br />

Silva, J.N.M., J.O.P. de Carvalho, J. do C.A. Lopes, B.F. de Almeida, D.H.M. Costa and L.C. de<br />

Oliveira. 1992. Growth and Yield <strong>of</strong>a Tropical Rain Forest <strong>of</strong>the Brazilian Amazon 13 Years<br />

After Logging. EMBRAPA and FCAP, Brazil.<br />

Sizer, N. 1992. Extractive Reserves in Amazonian; the Necessary Transformation <strong>of</strong>a Tradition. The<br />

Botany School, <strong>University</strong> <strong>of</strong> Cambridge.<br />

Smith, N.J.H., J.T. Williams, D.L. Plucknett, and J.P. Talbot. 1992. Tropical Forests and Their<br />

Crops. Ithaca, NY: Comell <strong>University</strong> Press.<br />

Stockdale, Mary, <strong>Oxford</strong> Forestry Institute. 1993. personal communication.<br />

Sullivan, F. 1990. The Fragile Forest. Timber Trades Journal, 9 June.<br />

61


62<br />

Swaine, M.D. and J.B. Hall. 1986. Forest Structure and dynamics. In: G.W. Lawson (ed.) Plant<br />

Ecology in West Africa. John Wiley & Sons Ltd.<br />

Swaine, M.D. and J.B. Hall. 1983. Early succession on cleared forest land in Ghana. British<br />

Ecological Society, London.<br />

Swanson, T.M. and E.B. Barbier. 1992. Economics/or the Wilds; Wildlife, Wildlands, Diversity and<br />

Development. London: Earthscan Publications Ltd.<br />

Terborgh, I. 1988. The big things that run the world - a sequel to E.O. Wilson. Conservation Biology<br />

2: 402-403.<br />

Thang, H. C. 1987. Forest management systems for tropical high forest, with special reference to<br />

Peninsular Malaysia. Forest Ecology and Management 21: 3-20.<br />

Toledo, V.M., A.!. Batis, R. Becerra, E. Martinez, and C.H. Ramos. 1992. Products from the tropical<br />

rain forests <strong>of</strong> Mexico: an ethnoecological approach. In: M. Plotkin and L. Famolare (eds.).<br />

Sustainable Harvest and Marketing <strong>of</strong>Rain Forest Products. Washington, D.C.: Island Press.<br />

Uhl, C., K. Clark, H. Clark and P. Murphy. 1981. Early plant succession after cutting and burning<br />

in the upper Rio Negro region <strong>of</strong> the Amazon basin. Journal <strong>of</strong> Ecology 69: 631-649.<br />

Uhl, C., H. Clark and K. Clark. Successional patterns associated with slash-and-bum agriculture in<br />

the upper Rio Negro region <strong>of</strong> the Amazon basin. Biotropica 14(4): 249-254.<br />

Uhl, C. and I.C.G. Vieira. 1989. Ecological Impacts <strong>of</strong> Selective Logging in the Brazilian Amazon:<br />

A Case Study from the Paragominas Region <strong>of</strong> the State <strong>of</strong> Para. Biotropica 21(2): 98-106.<br />

Vasquez, R. and A.H. Gentry. 1989. Use and misuse <strong>of</strong> forest-harvested fruits in the Iquitos area.<br />

Conservation Biology 3(4):350-361.<br />

Vincent, I.R., J.L. Chamberlain and S.T. Warren. 1987. Summary. In: F. Mergen and I.R. Vincent<br />

(eds.). Natural Management o/Tropical Moist Forests; Silvicultural and Management Prospects<br />

0/ Sustained Utilization. Yale <strong>University</strong>, School <strong>of</strong> Forestry and Environmental Studies, New<br />

Haven, cr.<br />

Wadsworth, F.R. 1983. Management Under Sound Ecological Principles. In: F. Mergen (ed.)<br />

Tropical Forests: Utilization and Conservation. Proceedings <strong>of</strong>an International Symposium held<br />

at Yale <strong>University</strong>, New Haven, CT.<br />

Wadsworth, F.R. 1987. Applicability <strong>of</strong> Asian and African silviculture systems to naturally<br />

regenerated forests <strong>of</strong> the Neotropics. In: F. Mergen and J.R. Vincent (eds.). Natural<br />

Management ojTropical Moist Forests; Silvicultural and Management Prospects <strong>of</strong>Sustained<br />

Utilization. Yale <strong>University</strong>, School <strong>of</strong> Forestry and Environmental Studies, New Haven, CT.


Wamulwange, G.L. 1993. Traditional Management Systems <strong>of</strong> Natural Resources: A Case for<br />

Western Province (Barotseland) Zambia. Paper presented at the Symposium for Indigenous<br />

People, 28-30 July, 1993. The Pueblo <strong>of</strong> Zuni, New Mexico.<br />

Wellner, P. and E. Dickey. 1991. The Wood Users Guide. The Rainforest Action Network, San<br />

Francisco, CA.<br />

Westoby, J. 1987. The Purpose <strong>of</strong> Forests; Follies <strong>of</strong>Development. <strong>Oxford</strong>: Basil Blackwell Ltd.<br />

Whitmore, T.C. 1991. Tropical Rain Forest Dynamics and its Implications for Management. In: A.<br />

Gomez-Pompa, T.C. Whitmore and M. Hadley (eds.) Rain Forest Regeneration and<br />

Management. Paris: UNESCO and The Parthenon Publishing Group Limited.<br />

Whitmore, T.C. 1992. An Introduction to Tropical Rain Forests. <strong>Oxford</strong>: Clarendon Press.<br />

Whitmore, T.e. and J.A. Sayer (eds.). 1992. Tropical Deforestation and Species Extinction. London:<br />

Chapman and Hall.<br />

Wilkie, D.S., J.G. Sidle and G.e. Boundzanga. 1992. Mechanized logging, market hunting, and a<br />

bank loan in Congo. Conservation Biology 6(4): 570-580.<br />

World Resources Institute, The World Conservation Union,and United Nations Environment<br />

Programme. 1992. Global Biodiversity Strategy; Guidelines for Action to Save, Study and Use<br />

Earth's Biotic Wealth Sustainably and Equitably.<br />

Wyatt-Smith, J. 1987. Problems and prospects for natural management <strong>of</strong> tropical moist forests. In:<br />

F. Mergen and J.R. Vincent (eds.). Natural Management o/Tropical Moist Forests; Silvicultural<br />

and Management Prospects <strong>of</strong> Sustained Utilization. Yale <strong>University</strong>, School <strong>of</strong> Forestry and<br />

Environmental Studies, New Haven, CT.<br />

Young, S. 1991. How plants fight back. New Scientist, 1 June: 41-45.<br />

Zerner, C. 1992. Indigenous Forest-Dwelling Communities in Indonesia's Outer Islands: Livelihood,<br />

Rights, and Environmental Management Institutions in the Era <strong>of</strong>Industrial Forest Exploitation.<br />

Report commissioned by the World Bank in preparation for the Forestry Sector Review.<br />

Ziffer, K. 1992. The Tagua Initiative: building the market for a rainforest product. In: M. Plotkin and<br />

L. Famolare (eds.). Sustainable Harvest and Marketing <strong>of</strong>Rain Forest Products. Washington,<br />

D.C.: Island Press.<br />

63

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!