You are on page 1of 24

Biodivers Conserv (2013) 22:301–324

DOI 10.1007/s10531-012-0429-5

REVIEW PAPER

What is the relevance of smallholders’ agroforestry


systems for conserving tropical tree species and genetic
diversity in circa situm, in situ and ex situ settings?
A review

Ian K. Dawson • Manuel R. Guariguata • Judy Loo • John C. Weber •

Ard Lengkeek • David Bush • Jonathan Cornelius • Luigi Guarino •


Roeland Kindt • Calleb Orwa • Joanne Russell • Ramni Jamnadass

Received: 30 August 2012 / Accepted: 18 December 2012 / Published online: 9 January 2013
 Springer Science+Business Media Dordrecht 2013

Abstract Smallholders’ agroforests may be valuable for conserving tropical trees


through three main mechanisms. First, trees planted and/or retained by farmers in agri-
cultural landscapes where wild stands were once found may be circa situm reservoirs of
biodiversity. Second, farmland trees may support conservation in situ by providing an
alternative source of product to reduce extraction from forest, and by acting as ‘corridors’
or ‘stepping stones’ that connect fragmented wild stands. Third, the additional value that

I. K. Dawson (&)  R. Kindt  C. Orwa  R. Jamnadass


The World Agroforestry Centre, Headquarters, P.O. Box 30677, Nairobi, Kenya
e-mail: iankdawson@aol.com

M. R. Guariguata
Center for International Forestry Research, P.O. Box 0113, BOCBD, Bogor 16000, Indonesia

J. Loo
Bioversity International, Via dei Tre Denari, 472a, Maccarese, 00057 Rome, Italy

J. C. Weber
The World Agroforestry Centre, West and Central Africa/Sahel Regional Office, BP E5118 Bamako,
Mali

A. Lengkeek
The Tree Domestication Team, Agro-business Park 76, 6708 PW Wageningen, The Netherlands

D. Bush
CSIRO Plant Industry, Black Mountain Laboratories, Clunies Ross Street, Black Mountain ACT 2601,
Australia

J. Cornelius
The World Agroforestry Centre, Latin America Office, c/o CIP, Apartado 1558, La Molina, Lima 12,
Peru

L. Guarino
Global Crop Diversity Trust, c/o FAO, Viale delle Terme di Caracalla, 00153 Rome, Italy

J. Russell
The James Hutton Institute, Invergowrie, Dundee, Scotland, UK

123
302 Biodivers Conserv (2013) 22:301–324

planting assigns to trees may result in greater interest in including them in seed collections,
field trials and field ‘genebanks’ that support ex situ conservation. Here, we critically
review the evidence for these mechanisms, and highlight areas for research and for
intervention so that agroforestry practices can better support conservation in each setting,
with an emphasis on often neglected genetic-level considerations. Based on current global
challenges to diversity, conservation will need to rely increasingly on a smallholder-farm
circa situm approach, but concerns on long-term effectiveness need to be properly quan-
tified and addressed. Connectivity between widely dispersed, low density trees in agri-
cultural landscapes is an important factor determining the success of the circa situm
approach, while improving farmers’ access to a diversity of tree germplasm that they are
interested in planting is required. The circumstances in which agroforestry plantings can
support in situ conservation need to be better defined, and research on the stability of active
tree seed collections (how long are species and populations retained in them?) as ex situ
reservoirs of biodiversity is needed.

Keywords Agroforestry  Circa situm conservation  Ex situ conservation 


In situ conservation  Effective population size  Genetic variation

Introduction

At the end of the last millennium, Oldfield et al. (1998) estimated that around 10 % of the
world’s 60,000–100,000 tree species were threatened with extinction. Since then, contin-
ued deforestation and the degradation of remaining forests and woodlands have led to
further deterioration of the situation (FAO 2010a). Concerns are reflected in the current
IUCN Red List of Threatened Species (www.iucnredlist.org), which indicates approxi-
mately 1,200 trees and shrubs as ‘critically endangered’, 1,700 as ‘endangered’, and
another 3,700 as ‘vulnerable’. Loss of forests and woodlands is a practical concern for rural
communities in the tropics that have traditionally depended on them for various resources
(FAO 2010a). Agroforestry practices that integrate trees with annual crop cultivation,
livestock production and other farm activities have received increasing attention as an
alternative to support livelihoods and ecosystem functions (Garrity 2004). ‘Land sharing’
involving crops, animals and trees can increase the efficiency of production systems and
support ecological and social resilience when different components occupy complementary
niches and react differently to disturbances (Kindt 2002; Kirschenmann 2007; Steffan-
Dewenter et al. 2007). Trees in farmland and in neighbouring natural forest fragments,
when present, also support populations of the insects, birds and other organisms needed for
crop pollination and biological pest control, adding to crop productivity (Ricketts et al.
2004; Harvey and Villalobos 2007; Tscharntke et al. 2008). Currently, more than 1.2
billion people practice agroforestry (www.worldagroforestrycentre.org; Leakey 2010) and
around 560 million people live in farm landscapes with 10 % or more tree cover (Zomer
et al. 2009).
Products realised from agroforestry trees include food, fodder, fuel, timber and medi-
cines (animal and human) for local use and sale, while services include soil fertility
replenishment, shade, water catchment protection and the fixation of carbon (Nair et al.
2009; Faye et al. 2011; Place et al. 2011). Agroforests are also considered important for
tree conservation (Scherr and McNeely 2008) and it is this role that we examine critically
in this essay, where we especially consider genetic (intra-specific) variation that is
important for tree function, but which has often not been taken into account properly in the

123
Biodivers Conserv (2013) 22:301–324 303

development of conservation principles and practices (Laikre et al. 2010). We here con-
sider the links between agroforestry and tree conservation in three different settings, circa
situm, in situ and ex situ, defined for the purposes of our discussion as follows:
• Circa situm conservation1 is the preservation of planted and/or remnant trees and
wildings in farmland where natural forest or woodland containing the same trees was
once found, but this has been lost or modified significantly through agricultural
expansion. This is the most obvious setting to consider for the role of agroforestry in
conservation, as the practice and the function are geographically coincident.
• In situ conservation is the maintenance of tree diversity in essentially wild forest and
woodland populations. In the context of the concerns of this paper, what is required is
an understanding of the relationship between agroforestry practices in one location and
conservation in adjacent natural forest or woodland sites.
• Ex situ conservation is the preservation of trees in seed storage, in field trials and in
field ‘genebanks’, and/or maintained in other ‘exotic’ locations outside the usual
environments, systems and/or geographic settings of the tree in question. Here, we are
primarily concerned with the possibilities for ex situ conservation that arise from
research on and/or wide distribution of a tree because of its value for smallholder
growers.
The boundaries between these settings are often in flux and can only be loosely defined.
Forests that at first glance appear pristine, for example, may have been manipulated for
millennia to produce ‘anthropogenic forests’. Such is the case in the Amazon, where
residual effects of Precolumbian human management are evident as high density aggre-
gations of useful trees such as Brazil nut (Bertholletia excelsa) located close to anthro-
pogenic soils known as ‘dark earths’ (Clement 1999; McNeely 2004; Clement and
Junqueira 2010; see also Sheil et al. 2012 for dark earths in Asian forest). Wiersum (1997)
proposed the term ‘‘co-domestication of forests and trees’’ to describe such effects. Again,
boundaries may be blurred when ancient forest cutting has reduced the natural range of a
species such that the division between its natural and exotic (through planting) distribution
is now unclear. The circa situm, in situ, ex situ division is however helpful for an
exploration of biodiversity threats and opportunities in the context of agroforestry prac-
tices, and in understanding which groups of trees species can be best conserved by different
methods in single or coordinated approaches. Research on this topic was identified as a
priority during recent consultations, especially in areas of the tropics with high biodiversity
and significant poverty, where interventions must be placed in the context of the benefits
local communities receive (CGIAR Consortium 2011). Here, by focusing on smallholders’
agroforestry systems, we demonstrate the importance of taking a wide view of conserva-
tion approaches in order to know where best to intervene for maximum impact and in order
to understand the unintended consequences of action in one setting on diversity in other
settings.

1
Also referred to in the literature as circa situ conservation; circa situm is formally the more correct term
(Heywood and Dulloo 2005) and is applied in the current paper.

123
304 Biodivers Conserv (2013) 22:301–324

The role of agroforestry in circa situm conservation

Tree species diversity in farmland can be high, but may be transitory

A range of case studies indicates that dozens and sometimes hundreds of tree species can
be found in agroforestry landscapes in the tropics (Table 1). Inventories indicate that, in
addition to planted trees, agroforests often contain many remnants from natural forest. In
some cases, these trees have been retained by farmers for the products they provide on an
ongoing (e.g., fruit, medicine) or one-off (e.g., large-bole timber) basis and/or for services
such as soil fertility replenishment or shade [e.g., shade coffee (Coffea spp.) and shade
cocoa (Theobroma cacao) production systems; see Table 1]. In other cases, trees have been
retained for cultural, religious (e.g., beliefs of the sanctity of the natural world) and/or
aesthetic reasons, and/or because although they are not used they are not overly compet-
itive and so there is no particular reason to remove them (a form of ‘benign neglect’;
Schroth and da Mota 2004; Schroth et al. 2011). The high tree species richness found in
these agroforestry landscapes suggests a strong role for farms in circa situm conservation.
Opportunities for conservation are supported by the positive role that tree species diversity
has been shown to sometimes play in increasing farm yields and promoting resilience
(Kindt 2002; Kirschenmann 2007; Steffan-Dewenter et al. 2007). In Niger, for example,
farmers explain that increasing the number of tree species per function minimises the risk
of ‘function failure’ in their farming systems, such that at least some species will provide
each required function even in the driest years (Faye et al. 2011).
Current levels of diversity in farmland may however be misleading for what can be
expected in future years. Functional diversity is more important than high tree species
numbers and ‘redundancy’ in this context means a significant proportion of diversity in
agricultural landscapes could be lost without having much effect on farm production, at
least in the short- to medium-term (Kindt et al. 2006). Shade cocoa production systems
(Table 1) illustrate transitions to lower and less valuable (from a conservation perspective)
farmland tree species diversity over time. For example, Rolim and Chiarello (2004) found
that late successional trees are becoming increasingly rare and early secondary succes-
sional species more dominant in the cabruca farms of northern Espı́rito Santo, Brazil,
leading the authors to question the long-term conservation value of these agroforests as
currently managed. Again, Sambuichi and Haridasan (2007) found reduced tree species
richness and/or a lower proportion of late successional trees and a higher proportion of
exotics in old compared to new cabruca plantations in southern Bahia, Brazil, suggesting
to the authors a loss in valuable diversity over time. The Kenyan surveys of Lengkeek et al.
(2005b) and Kehlenbeck et al. (2011) (Table 1) also illustrate that species-rich agroforests
can become dominated in overall abundance terms by exotics that are of limited conser-
vation concern. Clearly, the prospects for conserving tree species diversity in farmland
depend on the particular production system. While shade coffee and shade cacao culti-
vation are, for example, generally supportive of diversity, oil palm (Elaeis guineensis) and
‘full sun’ cacao cultivation are not; rather, the last systems are associated with a trend to
monoculture that reduces conservation value and encourages an unhealthy dependence by
growers on single crops (Donald 2004; Danielsen et al. 2009; Schroth et al. 2011; Ruf
2011; Waldron et al. 2012). Similarly, although rubber (Hevea brasiliensis) agroforests are
compatible with habitat and biodiversity conservation (Beukema et al. 2007), they are
often under pressure for conversion to more intensive and less diverse production systems
(Ekadinata and Vincent 2011).

123
Biodivers Conserv (2013) 22:301–324 305

Table 1 Examples of high tree species richness in tropical agroforestry landscapes (ordered by number of
species identified in each study)
Farming system and location Description of results References

265 Farm plots (each 0.5 ha) in 18 different 424 Woody plant species, 306 indigenous. Kehlenbeck
agro-ecological zones around Mount Mean of 17 species per plot. Eight of the et al. (2011)
Kenya, Kenya 10 most frequent species were exotic
35 Smallholders’ farms (60 ha in total) east 297 Tree species, *2/3 of which Lengkeek et al.
of Mount Kenya, Kenya indigenous. Mean of 54 species per farm. (2005b)
The five most common species were
exotic
5 Cacao cabruca plantation plots (each 293 Tree species, 97 % indigenous. Mean Sambuichi and
3 ha) in southern Bahia, Brazila of 101 species per plot. Exotic species Haridasan
were relatively more abundant (more (2007)
individuals per species) than indigenous
ones
146 Plots (each 0.063 ha) in 60 cacao 206 Mostly indigenous tree species. Mean Sonwa et al.
agroforests from 12 villages in the sub- of 21 species per agroforest. High relative (2007)
regions of Yaoundé, Mbalmayo and abundance of non-primary forest species
Ebolowa, Cameroon
24 Dairy farm pastures’ (237 ha in total) 190 Tree species, 57 % primary forest Harvey and
near Monteverde, Costa Rica trees. Primary forest trees accounted for Haber (2008)
(only) 33 % of all individuals
51 Plots (each 0.1 ha) in three shade coffee 123 Tree species identified (another 46 not Méndez et al.
cooperatives in Tacuba, El Salvador determined). Mean of 12–22 species per (2007)
plot, depending on cooperative. 11
species of conservation concern based on
international listings. Of 58 species
considered of benefit by farmers, 7 of
conservation concern
6 Forest gardens (2.68 ha in total) in two [120 Identified tree species (precise Marjokorpi
areas of West Kalimantan, Indonesia number not given). Mean of 52 identified and
species per garden. Most species in Ruokolainen
gardens not planted; of these, ‘easily (2003)
dispersed’ and/or ‘easily established’
species were over-represented in gardens
compared to forest
124 Plots (each 0.12 ha) in 15 shade coffee 107 Tree species, 83 indigenous (50 López-Gómez
farms of 3 types (shade monoculture, SM; primary and 33 secondary species). Mean et al. (2008)
simple polyculture, SP; diverse of 11 (SM) to 29 species (DP) per farm. 3
polyculture, DP) in central Veracruz, species of international conservation
Mexico concern. DP farms richer in tree species
than nearby forest
80 Plots (each 0.06 ha) in 20 cacao cabruca 105 Tree species, 101 indigenous, the Rolim and
farms in northern Espı́rito Santo, Brazila majority pioneer and early secondary Chiarello
species. Mean of 15 tree species per farm (2004)
Interviews with 68 cattle ranchers and 99 Tree species were identified by farmers Garen et al.
small-scale farmers in Los Santos and that they utilised, planted, or protected on (2011)
Rio Hato, Panama (NB, not direct farm their land, 3/4 of which indigenous
inventory)
60 Actively managed coffee-based 94 Species of mature trees, compared to Correia et al.
agroforestry plots (of variable area) in 3 134 in natural forest. Mean of 59 tree (2010)
villages in Guinée Forestière, Guinea, species per village. A few species
West Africa dominant in agroforests. Nine species in
agroforests classified as vulnerable
according to IUCN listings

123
306 Biodivers Conserv (2013) 22:301–324

Table 1 continued

Farming system and location Description of results References

240 Plots (each 0.2 ha) in coffee farms in 92 Identified trees species in coffee plots Philpott et al.
Lampung province, Sumatra, Indonesia, outside BBS, 90 in plots inside of BBS, (2008)
outside and inside (120 plots each) Bukit compared to 141 in natural forest plots
Barisan Selatan National Park (BBS) (with same sample area). The most
abundant species in coffee plots outside
and inside BBS were exotic
a
Cabruca is a type of agroforestry system in which cacao is planted in cleared understorey within native
forest

A point of equal concern that may limit the long-term value of agricultural landscapes
for conserving trees is the very low densities of many species. Of the 107 tree species
reported in coffee agroforests in Veracruz, Mexico, for example, 71 % were found in only
one or two of the 15 farms surveyed (López-Gómez et al. 2008), while in coffee farms in
Guinea many forest trees were represented by only a few individuals (Correia et al. 2010).
Again, in cabrucas in Brazil, of the 105 tree species recorded by Rolim and Chiarello
(2004), 43 occurred only once, while 80 occurred no more than five times, in a total sample
area of 4.8 hectares. Similarly, of the 297 tree species identified by Lengkeek et al. (2005b)
in Central Kenyan farms, more than 40 % occurred at a density of less than 0.1 mature
trees per hectare (averaged over the total surveyed area of 60 ha). Typical of other surveys,
most of the low density species identified by Lengkeek et al. (2005b) were indigenous in
origin and many were remnants.
The reasons why low densities are a concern are two-fold. First, they make individual
species vulnerable to the decision making processes of particular farmers on whether to keep
or cut a tree: one farmer deciding to cut could entirely remove a species from the landscape.
Second, low densities have implications for regeneration when species are outcrossing (as
most trees are; Petit and Hampe 2006). Trees can sometimes pollinate each other over very
long distances in agricultural landscapes and in farm-forest mosaics (see below), but some
impairment of pollination (in frequency and reliability) can be expected for a proportion of
farm trees occurring at very low densities (Lowe et al. 2005; Ward et al. 2005). This may be
worst for animal-pollinated trees where anthropogenic climate change reduces pollinator
abundance and introduces pollinator-tree life cycle asynchronies (FAO 2008; Gilman et al.
2012). If rare trees are unable to pollinate each other and therefore do not set seed, from a
regenerational perspective they are ‘the living dead’ sensu Janzen (1986).
At what point low density begins to significantly restrict pollination in farmland is
unknown for most trees, and will depend on the particular pollinator or pollinators, and the
specific farm landscape and/or farm-forest mosaic. However, returning to the example of
Lengkeek et al. (2005b), we can perhaps surmise that the prospects for long-term circa
situm conservation of the small proportion of relatively high density indigenous tree
species (\10 % at [5 mature individuals per ha) found in farmland are good, while the
prospects for the high proportion of species at very low density (40 % at \0.1 mature
individuals per ha) are poor. In the long-term, these last species may be more effectively
conserved in situ or ex situ rather than wasting effort and capital on circa situm approaches
(although see below on the inter-relatedness between circa situm and in situ). The key
point is that reproductively inactive trees in agricultural landscapes, perhaps present as a
result of relatively recent (in generational terms) forest clearance that has left a few
remnants only, may lead to complacency about the long-term tree species carrying capacity

123
Biodivers Conserv (2013) 22:301–324 307

of farmland. Research is required on densities and configurations needed for preserving a


range of trees with different reproductive biologies, managed by farmers in different
ways, and in different species assemblages, in a range of farmland environments. One
simple approach is to observe which trees are producing seed in the landscape and then
test this seed in nursery experiments to determine if the germplasm is viable and pro-
duces vigorous seedlings. Observations on pollinator identities and their characteristics
are also required.
Finally, the density of a particular tree in farmland depends at least in part on how it is
used. Considering function is therefore important when exploring conservation options, as
demonstrated by the examples described in Table 2. Species used for soil fertility
improvement or fodder provision, for example, are likely to be found at higher densities in
farm landscapes than those used to provide (locally used only) medicines, so circa situm
conservation may be easier for trees with the former functions (though they may ‘push out’
trees with other functions). Unfortunately, in several regions of the tropics the provision of
services such as soil fertility improvement and products such as fodder often relies cur-
rently on exotic trees that are or little interest from a conservation perspective, but greater
emphasis could in future be placed on indigenous species (ideally, a mix of indigenous
species).
In summary, decisions about the long-term value of agroforests for circa situm con-
servation of tree species diversity should clearly not be based on current levels of species
diversity alone. Rather, the occurrence and abundance of exotic trees, transitions in
diversity that are determined by regenerational behaviour and density-dependent repro-
ductive biology, and methods of production and market development, are all important
factors in determining value. The species carrying capacity of farmland needs to be
observed over time in specific locations in order to identify the major factors determining
change.

Genetic diversity in farmland tree populations appears generally high, despite concerns
to the contrary

Many outbreeding trees are susceptible to inbreeding depression that is manifested in


many ways, including through low fruit set and low vigour, poor growth form and early
mortality of progeny (Charlesworth and Charlesworth 1987). From a circa situm con-
servation perspective, resultant losses in production are important because they are a
disincentive to farmers to grow and maintain trees (Dawson et al. 2009 and references
therein). Inbreeding depression is prevented when a group of trees maintains a large
enough effective population size (Ne), which is a function of the genetic diversity a
stand contains and the connectivity (gene flow) between trees in the landscape (Lowe
et al. 2005). Concerns about possible low tree genetic diversity in agricultural land-
scapes, which could lead to inbreeding, have been expressed based on the narrow
collection practices (seed taken from only a few trees) that farmers and tree nursery
managers often adopt when they sample germplasm (Lengkeek et al. 2005a). Fears have
also arisen that ‘founder effects’ in farmland may be compounded during subsequent
regeneration, as once farmers have planted a tree they tend to turn to it for seed for
future planting rounds, rather than returning to ‘external’ sources (see Weber et al. 1997
for an example from the Peruvian Amazon, Lengkeek et al. 2005b for a Kenyan
illustration).
Molecular marker studies of farmland and adjacent forest tree populations have begun
to shed light on relative levels of genetic variation and show that diversity in agricultural

123
308 Biodivers Conserv (2013) 22:301–324

Table 2 Relationships between the functional uses of trees and abundance in farmland
Functional use Abundance Implications/possible interventions to
support circa situm conservation

Fruits and nuts Trees planted for their fruits and nuts will Uniquely, the desired product is the
for food probably occur at low- to medium-density propagule: to fulfil function, trees must be
in farmland when for subsistence use, pollinated. For outcrossing species, loss
high density if produce for commercial of landscape connectivity means the
sale raison d’être for farmers to plant and/or
retain trees is absent. Encourage farmers
to grow useful indigenous fruits in place
of exotics that are sometimes found in
greater abundance in farms
Timber Timber trees may be low density remnants Whether or not planted woodlot trees will
in farmland, allowed to stand when seed and contribute to the next generation
clearing natural forest. When planted in depends on the stage and method of
woodlots, densities are higher harvesting (e.g., harvesting of young
poles/coppicing may prevent seeding).
Do not cut a proportion of woodlot trees
in order that they seed. Focus on planting
useful indigenous species
Medicinals If for traditional use, only a few trees may Networks of traditional healers may focus
be required per farm/village and densities on retaining or establishing particular tree
may be very low. If commercial planting, species in specific locations for wider use,
densities may be high with exchange of material. For
commercial harvest, focus on planting
useful indigenous species
Soil fertility Improved planted fallow technologies Depending on the length of the rotation,
improvement usually involve establishment of trees or tree clearance and crop replanting may
shrubs at high density across the fallow occur before trees reach maturity. Retain
some trees so that they produce seed.
Exotics are widely used but indigenous
species can be productive and use should
be encouraged
Fodder Fodder trees are often planted in fence rows Lopping of trees for fodder may limit
at relatively high densities to provide seeding. Allow some trees to produce
sufficient feed for an animal from a single seed. Exotics are widely used but
small farm indigenous species often could be; focus
on the latter where productive
Shade Shade trees may be natural forest canopy Since canopy formation is important to
with crops planted below, with a wide provide shade, most trees will seed,
range of species for which individual contributing to the next generation. To
densities are low. Trees planted for shade support conservation, promote crop
will have higher densities of individual species and varieties that perform better
species under shade conditions

landscapes depends on the length and intensity of tree management, the degree of planting
undertaken and the method of propagation. When species have been managed for mil-
lennia, bottlenecks can be observed in farmland [e.g., Hollingsworth et al. 2005; Dawson
et al. 2008 for inga (Inga edulis) in the Peruvian Amazon; Miller and Schaal 2005, 2006 for
jocote (Spondias purpurea) in Mesoamerica, both important fruit trees; see also Miller and
Gross 2011 for further information on fruit tree domestication and impacts on diversity].
On the other hand, if the intensity of management is low, there can be little difference
between farm and neighbouring natural populations [e.g., Kelly et al. 2004 for sheanut
(Vitellaria paradoxa), an important fruit tree in the Sahelian and Sudanian eco-zones of

123
Biodivers Conserv (2013) 22:301–324 309

West Africa; Lengkeek et al. 2006 for Meru oak (Vitex fischeri), a timber tree, in Kenya].
Overall, the limited molecular data available from comparative studies suggest that
although indigenous trees in farmland occasionally experience genetic bottlenecks com-
pared to natural stands, so far these have been relatively minor and do not appear to detract
significantly from on-farm stands having a conservation function, despite the already
expressed concerns connected with germplasm collection. Perhaps bottlenecks in planted
circa situm populations are limited because different smallholders initially obtain germ-
plasm from different wild source trees; with only small farms, landscape-level diversity is
thereby retained through source mixing. Likely, the problem will come if and when the
number of possible source trees in the landscape is greatly reduced (because they are more
likely to be introduced from fewer sources, founder effects and impacts on production may
be of more concern for exotic trees than indigenous ones; see, e.g., Harwood et al. 1992,
1997 for the case of the timber tree Grevillea robusta introduced from Australia into
Africa). It seems likely that of most concern presently for the Ne of indigenous trees in
farmland is not low genetic diversity but low overall abundance and limited inter-tree
connectivity (see previous section).

The role of agroforestry for in situ conservation

Although it is often claimed to do so, tree cultivation may not decrease harvesting
pressure on natural stands

It has been widely proposed that cultivating trees in smallholdings around natural forests
and woodlands is a means to conserve wild stands by providing an alternative resource for
harvesting (e.g., Dawson 1997; Lambert et al. 1997; Strandby-Andersen et al. 2008;
Jamnadass et al. 2010). The example of Murniati et al. (2001), based on research on the
island of Sumatra in Indonesia, is often quoted. There, smallholders that planted agro-
forestry gardens and rice plots extracted fewer resources from natural forest than farmers
that only practised rice cultivation, presumably because the latter had only limited access
to tree products on their farms and so had to obtain them from the wild. The same principle
has been posited for large-scale plantation establishment, with the example most com-
monly quoted being commercial forestry in New Zealand (Brockerhoff et al. 2008). There,
the establishment of fast growing exotic timber trees in plantations was reported by the
above authors to result in the spatial separation of forests’ production and conservation
functions, an achievement facilitated by the proper coordination of interventions and the
ability to control harvesting practices.
Although at first glance intuitive, a link between tree cultivation and a reduction in wild
harvesting should not be taken for granted (Newton 2008) and the conventional wisdom
that ‘‘plantation forestry [and agroforestry] can reduce pressure on natural forests’’ is
largely anecdotal (Paquette and Messier 2010). There are several reasons why there may
not be a positive link. One is that planting may result in less priority being placed on the
sustainable management of natural stands. Clapp (2001), for example, reported that in
Chile the development of timber plantations from the 1960s resulted in foresters viewing
natural timber stands merely as ‘stopgap’ supplies, a resource that could be over exploited
while plantations matured. Another reason is that cultivation may stimulate the develop-
ment of markets and infrastructure that unintentionally ‘capture’ wild resources as well as
serving, as intended, the harvest of planted stands (Cossalter and Pye-Smith 2003; An-
gelsen and Kaimowitz 2004; Marshall et al. 2006). A third reason is that tree planting when

123
310 Biodivers Conserv (2013) 22:301–324

it leads to a profitable business may trigger forest and woodland clearance for further
cultivation. In this way, planting of a tree of local origin in previously forested land cleared
for this purpose may promote circa situm conservation of the species in question while at
the same time hindering in situ preservation of the tree and a whole range of other taxa
(Dawson and Jamnadass 2007). The expansion of oil palm plantations in Indonesia is an
example that is doubly bad for biodiversity: not only is the species richness found within
plantations low compared to other woody perennial production systems, but there is a
demonstrable link to the cutting of extra forest to provide the land required; furthermore,
the varieties of oil palm planted are themselves generally of little or no conservation value
since they are exotic and primarily widespread commercial types (Donald 2004; Danielsen
et al. 2009). Although some authors have suggested that intensive cultivation reduces the
amount of land needed for production and hence the amount of forest that need be cut,
there are few quantitative data to support the view that ‘land sparing’ in this way is more
effective than ‘land sharing’ as a conservation approach (Balmford et al. 2012; Tscharntke
et al. 2012). ‘Land sparing’ is for example unlikely to prove a viable strategy when the
market demand for a product is far from fulfilled, as is the case for palm oil. Even if
intensification of production results in greater total land area protected in situ from agri-
culture, there are clearly no guarantees that these areas will effectively conserve trees
under other forest use pressures (Hayes 2006).
The literature that is available on medicinal plants, among which many trees are
numbered, is also cautionary on the in situ conservation benefits that can be expected from
the cultivation of a particular tree (Homma 1996; Schippmann et al. 2002; Marshall et al.
2006; Wiersum et al. 2006; Newton 2008). Generally, concerted planting of medicinals
begins only after wild resources have already been over harvested, which may be akin to
closing the stable door after the horse has bolted from an in situ conservation perspective,
although local cultivation can of course still support circa situm conservation in these
circumstances. Certainly, current data suggest that tree planting in agroforestry systems is
not necessarily supportive of wild stands: the overall effect may be positive, neutral or
unfavourable. Detailed research to establish when and where positive results can be rea-
lised is required, based on trees such as pygeum (Prunus africana, a medicinal tree) and
allanblackia (Allanblackia spp., providing edible oil) (Box 1). Both of these trees are
currently undergoing domestication, have local and international markets, and are being
grown in smallholdings surrounding forests where they are found naturally. Market
demands for product traceability, sustainability and uniformity may be factors that promote
beneficial linkages between cultivation and in situ conservation, and certification schemes
that confirm location of production, germplasm source and farming method may be
important (Strandby-Andersen et al. 2008).

Agroforestry trees may promote connectivity between natural forest fragments,


but the results may not always be positive

Another way that farmland trees may support wild tree stands and natural populations of
other flora and fauna is by acting as ‘corridors’ or ‘stepping stones’ for seed dispersers,
pollinators and other migrating animals between otherwise isolated natural forest frag-
ments, helping to maintain Ne above minimum critical levels for survival (Bhagwat et al.
2008; Doerr et al. 2010; Gilbert-Norton et al. 2010). Pollen-mediated gene flow between
trees in forest and farmland has been studied using molecular markers and paternity
exclusion analysis techniques (Ward et al. 2005; Jha and Dick 2010). As expected, these
studies demonstrate that pollen exchange is generally more frequent between nearby trees;

123
Biodivers Conserv (2013) 22:301–324 311

Box 1 Can cultivation in agroforests preserve trees in natural forests? Examples for research
Two interesting case studies where exploration of linkages between cultivation and in situ conservation is
merited are the African trees pygeum (Prunus africana) and allanblackia (Allanblackia species), both of
which are currently being planted by African farmers.
Pygeum The bark of pygeum, the only species of the cherry genus that grows naturally in Africa, provides an
extract used worldwide to treat benign prostatic hyperplasia (Simons et al. 1998). The tree is found in
montane forests across Africa, but wide scale harvest from natural stands in Cameroon and Madagascar in
particular resulted in it being listed in 1995 under Appendix II of the Convention on International Trade in
Endangered Species of Wild Fauna and Flora (CITES). Since listing, periods of export boom have been
followed by bust as voluntary suspensions and outright bans to counter over exploitation have at times
been adopted. Over the last two decades, the cultivation of pygeum has been promoted in Cameroon to
provide an alternative source of bark and to improve local peoples’ livelihoods (Dawson et al. 1997; Gyau
et al. 2012). Thousands of smallholders have planted trees that are now reaching harvestable age, and a
moratorium on bark export from Cameroon was recently lifted. Research is required to determine to what
degree bark from cultivated stands is able to substitute for that from natural populations, and how this can
best be done (e.g., through the formation of farmer producer groups to coordinate farm sales?). In addition,
if a switch to farm harvesting does occur, what impact will this have on the livelihoods of the collectors of
wild bark who are often among the poorest in communities and do not have access to land to plant trees
on? And, what would be the effect of collection from farms on the attitudes of farmers and wild harvesters
to forest management?
Allanblackia The seed of allanblackia, a genus of nine species native to biodiversity hotspots in the humid
forests of sub-Saharan Africa, yield edible oil with significant global market potential. Within the last
decade, wild harvest for export has begun in Ghana and Tanzania, and cultivation to improve
smallholders’ incomes has commenced (Jamnadass et al. 2010). Unlike many other non-timber forest
products, cultivation has begun at the same time as market development, which may help to reduce wild
over exploitation and associated negative impacts on biodiversity (Dawson and Jamnadass 2007). The
worst case scenario would however involve wide scale cutting of forest for monoculture establishment. An
emphasis on production by smallholders—who generally prefer to plant multiple crops and rely on natural
forests for a range of products—is one approach to protect farmland and forest biodiversity, but
effectiveness needs to be tested. Large productivity gains appear to be possible under cultivation and this
may afford greater protection for allanblackia in the wild as collection of the latter will be relatively less
profitable (Peprah et al. 2009). Will early cultivation be effective in taking pressure off natural stands by
directing market demand for seed oil to planted sources? And, how committed are commercial, research
and conservation organisations to resolve the practical and reputational challenges involved in developing
a sustainable business? If the development of the allanblackia business model fails due to a lack of
profitability or environmental concerns, what will be the impact on the attitude of local communities to the
use and management of other indigenous trees and natural forest more widely?
Addressing the linkages between cultivation and wild harvesting is difficult as many political, social and
economic factors need to be considered, and decades may be required to collect data before firm
conclusions can be drawn. It is possible that ‘co-opting’ of the production of pygeum and allanblackia by
large commercial growers in plantations in locations outside the native ranges of the species could protect
the natural resource base if production is much more efficient (i.e., ex situ cultivation could then promote
in situ conservation), but of course local people would no longer benefit from harvesting and cultivation,
and this would counter development objectives.

transfers can however sometimes occur over large distances in landscape mosaics, with
even isolated farmland individuals sometimes able to contribute to maintaining Ne of trees
in fragmented natural forest (Boshier 2004; Kramer et al. 2008). An extreme example is
mahogany (Swietenia humilis, an important timber) in the deciduous dry forests of Hon-
duras, where White et al. (2002) found that a lone individual in pastureland participated in
insect-mediated pollen exchange with forest trees over distances of more than 4.5 km.
Clearly, as well as supporting tree conservation in situ, such pollen exchange should
support conservation circa situm when farm populations are small. Further research is

123
312 Biodivers Conserv (2013) 22:301–324

needed to determine tree planting configurations and species mixtures in smallholdings that
best support pollen-mediated gene flow between forest fragments.
When it is planted rather than remnant farmland trees that pollinate neighbouring wild
stands, however, the effects may not always be positive because of concerns regarding the
source of planted material. If the source is not local, then planted trees are likely to have a
different genetic composition from wild stands, and crossing between them could lead to
the dilution and loss of unique diversity in the wild, with the possible extinction of
populations, and outbreeding depression is possible through the breakdown of co-adapted
gene complexes (Ledig 1992). Jamnadass et al. (2005) raised reservations on this basis on
the (then) common practice of transferring sesbania seed (Sesbania sesban, a shrub planted
by smallholders to enhance soil fertility) from Kenya to Zambia (because Kenyan stands
were easier to collect seed from to supply to farmers), as the genetic compositions of
natural stands found in the two locations are very different. Similarly, wild coffee (Coffea
arabica) stands in the small fragments of Ethiopian montane forest that still exist are
threatened by hybridisation with coffee cultivars introduced from elsewhere into neigh-
bouring farms (Labouisse et al. 2008; Aerts et al. 2013). Again, planted inga trees in
smallholders’ fields in the Peruvian Amazon were found to be different in genetic com-
position from neighbouring wild populations, raising concerns for in situ stands if there are
interactions between forest and farmland individuals (Dawson et al. 2008). In this instance,
‘exotic populations’ of cultivated inga had larger fruit than local wild trees, explaining why
farmers had introduced them. In such circumstances, understanding the trade-offs between
connectivity, genetic dilution and possible outbreeding, and the level of return farmers
receive from planting, is necessary.
The key point is that promoting connectivity may not necessarily support tree popu-
lations in situ if farmland trees are of the ‘wrong’ source, chosen as such either inadver-
tently or because they are the most productive trees for farmers to plant. In some situations,
concerns have even been expressed that cultivated-wild tree hybridisation could result in
traits introduced into cultivars through genetic modification (GM) being transferred into
natural stands, with potentially significant evolutionary consequences in the wild [see
Delplancke et al. 2012 for concerns regarding cultivated (Prunus dulcis) and wild (Prunus
orientalis) almond]. If the impacts of hybridisation are considered likely to be highly
negative for wild stands, an alternative approach may be taken in which cultivated material
and natural populations are deliberately isolated; this would then be an argument for
planting exotic rather than indigenous trees around natural forests and woodlands in order
to support in situ conservation (Potts et al. 2001). However, exotic tree species may invade
forests, woodlands and other ecosystems through seed dispersal (Williams and Wardle
2009); indeed, the problems of invasiveness of exotic trees compared to the benefits
received from planting is the subject of lively debate (see, e.g., Low 2012a, b; Kull and
Tassin 2012 for opposing views). A potential solution to the hybridisation conundrum is
local selection of the best performing wild material for planting, thereby addressing source
and productivity concerns. One method of domestication that fulfils this requirement is the
participatory domestication approach (discussed below). Finally, although the orthodox
view is that wild-cultivated interactions may be detrimental to wild stands if cultivated
material is very different genetically, some scientists have advocated the mixing of dif-
ferent populations under scenarios of rapid anthropogenic climate change, where wild
stands perhaps need completely new combinations of alleles to adapt sufficiently quickly
(Weeks et al. 2011). Clearly, research is required on the magnitude of outbreeding
depression for model indigenous tree species of importance to conservationists, foresters
and smallholders, as it remains a relatively understudied phenomenon and there is limited

123
Biodivers Conserv (2013) 22:301–324 313

3,000
All sources Natural populations

2,500

Number of species 2,000

1,500

1,000

500

0
0 10 20 30 40 50 60
Number of suppliers

Fig. 1 The relationship between tree species (n = 5,874) and suppliers (n = 144) in the World
Agroforestry Centre’s Tree Seed Suppliers Directory. Data for all sources of seed (natural populations,
naturalised stands, farmers’ plantings, etc.) and natural sources only are shown. Many species have only
single suppliers, while for natural sources in particular only a low proportion of species (103 species in total)
have five or more suppliers. Low across supplier replication and limited natural sources indicate that
considering species numbers alone may give an unduly positive impression of the value of active seed
collections for ex situ conservation

hard evidence except in cases of interspecies hybridisation (Ellstrand 2003; Edmands


2007).

The role of agroforestry in ex situ conservation

‘Active’ seed collections and long-term seed ‘genebanks’ contain many tree species,
but their conservation value may be limited

Tropical trees with orthodox seed characteristics that are important to smallholders and
other growers have been widely collected to support research and planting initiatives, and
can be found stored in ‘active’ seed collections maintained by tree seed centres, scientific
institutions and private commercial suppliers. To provide a handle on the number of tree
species in such collections, we reviewed the World Agroforestry Centre’s open access Tree
Seed Suppliers Directory2 (the TSSD), which lists germplasm availability for 144 suppliers
located around the world. Our review indicated an impressive 5,874 woody perennial
species present in the collections of these suppliers and suggested that, in toto, they are an
important ex situ reservoir of biodiversity. First impressions may however be deceptive, as
shortcomings are evident on further examination of the TSSD to assess the following: (1)
the coverage of collections across institutions; and (2) the origin of the germplasm stored.
On the first point, only 19 % of listed species had five or more suppliers, while a high
proportion (49 %) had only a single supplier (Fig. 1). Generally, active collections have no
2
The TSSD, the only major global effort to catalogue tree seed suppliers, is available at www.world
agroforestry.org/Sites-old/TreeDBS/tssd/treessd.htm.

123
314 Biodivers Conserv (2013) 22:301–324

particular commitment to conservation and species with single or only a few suppliers are
therefore vulnerable to the vagaries of changing research priorities in the public sector
and shifts in the commercial imperatives of the private sector: the presence of such
species in active collections should therefore be considered precarious. On the second
point on origin, only about half the total number of species in the TSSD (2,846) had
known natural sources, with 68 % of these species being held by a single supplier; only
103 species (fewer than are often found growing circa situm in particular farm land-
scapes, see above) had known natural sources and were held by five or more suppliers.
The source of germplasm is important from a conservation perspective because when
origin is known material can be placed in its geographic context and population coverage
within species can be assessed; the fact that there are relatively few species of known
origin and with broad coverage across institutions is therefore a concern. Our analysis of
the TSSD therefore suggests that high overall species numbers may lead to a false sense
of security of the benefits of active tree seed collections for conservation. As the existing
TSSD is now more than a decade old, data are currently being updated for a revision
planned for release in 2014 (coordinated by RK and CO); comparison of old and new
versions will provide some information on the stability of suppliers and the collections
that are held by them.
Storage in seed ‘genebanks’ that are not dependent on current research or commercial
concerns for their survival and that use optimal conditions for long-term preservation may
be a more secure means of ex situ conservation. Such genebanks are common for agri-
cultural staples and interest in including other plants has grown (Schoen and Brown 2001).
The genebank with the most comprehensive coverage is the Millennium Seed Bank
(Kew, UK), which currently holds seed of over 10 % of the world’s wild plant species
and by 2020 aims to hold 25 %, with many trees included (www.kew.org/science-
conservation/save-seed-prosper/millennium-seed-bank). Although at first sight an attrac-
tive conservation method for orthodox-seeded trees, a significant limitation is the need to
regenerate stored samples periodically. This is expensive and difficult for trees due to long
generation intervals, large growth forms, generally outbreeding reproductive systems and
species (and occasionally population)-specific regeneration requirements, sometimes
including the need for particular microsymbionts (Petit and Hampe 2006). Costs are
especially high if a representative range of populations is to be maintained rather than just
one or a few stands of each species. Understandably, the approach to date has been to put
tree seed into long-term storage and worry about regeneration later. Seed genebanks may
however give a false impression of ‘available’ diversity, in the sense that this diversity can
be reintroduced into a wild or agricultural environment and used in the future in any
meaningful way. In addition, storage in seed genebanks prevents a key feature of in situ
and circa situm conservation which is dynamism, the potential for continued evolution
under natural and human selection processes (including anthropogenic climate change;
Groves et al. 2012). Undertaking regeneration with growers that obtain some benefits from
planting is a way of reducing land rental and maintenance charges that are the major costs
in the process, but in many cases costs are surely prohibitive. Research on approaches that
maintain viability in storage and therefore increase the time before seed need be regen-
erated, such as seed cryopreservation, is therefore critically important. However, seed
storage is not yet, and may never be, a viable option for most of the many tropical trees
with highly recalcitrant seed, including many fruit trees valued by smallholders (Schmidt
2007).

123
Biodivers Conserv (2013) 22:301–324 315

Field trials, field genebanks and other ‘live’ plantings are important
for the conservation of some trees, but may divert attention from more useful activities

For economically important trees with recalcitrant seed and/or that are propagated clonally
to maintain genetic integrity, ex situ field plantings play a dual role, first in evaluating
genetic variation to boost production and second in conservation. Tree commodity crops
such as cacao and coconut (Cocos nucifera) provide good examples. Large field collections
of cacao include the International Cacao Collection in Costa Rica and the International
Cocoa Genebank in Trinidad and Tobago, which consist of around 3,000 and 1,100
accessions, respectively (FAO 2010b). Both locations are nations where smallholder cocoa
production is important and the costs of maintaining collections are therefore justified
outside the native range of the species in western Amazonia. For coconut, a novel ex situ
approach to conservation termed polymotu (‘many islands’) has recently been proposed,
which builds on an ancient Polynesian practice in which single varieties are grown on
isolated islands to maintain their genetic integrity (Bourdeix et al. 2011).
For lesser-used tree species, costly maintenance may make field genebanks financially
unviable. Perhaps the worst case scenario is when resources that could be used to promote
the use and conservation of a species in farmland are instead squandered on developing
genebanks that are too costly to maintain over the long-term and so after a few years are
allowed to fall into disrepair; the result is that conservation benefits are lost and more
important work that could have benefited farmers was not undertaken (such a scenario has
been posited for various underutilised palms in South America; van Leeuwen et al. 2005).
One way to maintain live collections of more unusual tree species is in botanic gardens and
arboreta where the costs of maintenance can be linked to educational and public awareness
activities. In the past, such collections were however generally made to show plants to
visitors rather than for conservation; often not much attention was paid to germplasm
source, which is frequently unknown (but clearly narrow), which limits utility (Oldfield
2009; Kozlowski et al. 2012).
As discussed above, when considering agroforestry smallholdings as locations for tree
conservation, the concern is generally with indigenous species that are being preserved circa
situm. There are however occasions when tree species have been distributed outside their
native ranges and then through natural selection and conscious and/or unconscious human
manipulation in exotic locations have become landraces with unique adaptive and pro-
ductive characteristics that are important to growers. Examples include the tree commodity
crops such as cacao and coconut already mentioned, along with coffee and mango
(Mangifera indica) that have been widely dispersed and manipulated over centuries (Mohan
Jain and Priyadarshan 2009). In these cases, on-farm conservation in exotic locales may be
important for preserving landraces. It should be noted, however, that in general exotic trees
that are assumed to be landraces may in fact only occasionally outperform new introduc-
tions of germplasm from the native range; indeed, they may often underperform because
they were taken from inferior natural stands initially and/or have experienced bottlenecks
during distribution (Otegbeye 1991; Harwood et al. 1992, 1997; Varghese et al. 2009).
More research is required on the performance of so-called landraces relative to natural
stands, through comparative field trials, to determine which are worthy of conservation. If
apparent landraces perform badly, there appear to be few good reasons to try to conserve
them in farm landscapes if the same trees can still be found in their native environments
and/or in seed collections; from a conservation perspective, the focus is then better placed
on replacing ‘landraces’ in farmland with indigenous tree species. In a few extreme cases,
tree species are extinct in the wild and are found only under cultivation at exotic sites, as

123
316 Biodivers Conserv (2013) 22:301–324

observed for the toromiro tree (Sophora toromiro) in the Pacific (Maunder et al. 2000). In
such cases, ex situ conservation may be the only option. In more cases, species are not
extinct in the wild but particular populations are and are only maintained ex situ in
breeding programmes [e.g., for Vanuatu whitewood (Endospermum medullosum) in the
Pacific; Doran et al. 2012].

Practically promoting circa situm conservation

Anthropogenic climate change is resulting in locational shifts in environments where


particular forest assemblages, tree species and populations can survive and thrive, at a
speed too fast for many wild populations to respond through adaptation or by natural
migration to more physiologically appropriate sites, which in any case may already be
occupied by other land uses (Malcolm et al. 2002; Malhi et al. 2009). At the same time,
the current rate of global deforestation is around 13 million ha per annum, with Africa
and South America particularly affected recently, due in large part to the conversion of
forest to agricultural land to feed expanding human populations that also have changing
dietary preferences that require the use of more land per capita (FAO 2010a). For moist
tropical forests in particular, tree cutting, elevated temperatures, altered precipitation
levels and extreme weather events present significant threats to the conservation of tree
species and populations in situ (Giam et al. 2010; Peres et al. 2010). At the same time,
ex situ methods for conservation are constrained primarily by limited representation and
the problem of regeneration. It seems inevitable therefore that in the future there will be
a greater reliance on the conservation of trees circa situm in smallholders’ farm land-
scapes. How, then, can farmland tree species diversity be maintained or enhanced over
time or at least the rate of loss of diversity be reduced? How can the genetic base of
farmland tree populations be retained while supporting productivity increases that
encourage farmers to plant them? And how can circa situm conservation respond to
climate change?
Rolim and Chiarello (2004) suggested various interventions to improve the conservation
value of cacao agroforestry systems, including the eradication of non-native (especially
invasive) species in farmland and the protection of saplings of particularly important native
trees (especially late successional species; see also Cassano et al. 2009 for measures to
protect wider floral and faunal diversity in cabrucas). If remnant farmland trees are con-
nected with other trees in farmland or forest through pollination, then the protection of
natural regeneration from the seed crop underneath these trees is a valid strategy to support
their conservation. This approach may however not work for reproductively isolated
obligate outcrossers and in this case germplasm may need to be deliberately introduced.
Alternatively, the introduction of new and possibly alien pollinator populations may
enhance connections (Dick 2001). Similarly, remnant trees that act as perches for birds that
disperse seed can allow the regeneration of a range of plants underneath them with the
suitable protection of new growth; this is unless the distance between forest and farmland
trees is too great and/or farmland trees are too rare to encourage bird flights (Guevara and
Laborde 1993; Duncan and Chapman 1999; Holl 1999).
Measures to improve smallholders’ access to tree planting material and thereby enhance
diversity were reviewed by Dawson et al. (2009; see also references therein) and include
‘diversity fairs’, the training of seed collectors in appropriate germplasm sampling methods
for a wide range of tree species, the provision of more business support to small-scale
commercial tree seed and seedling enterprises that may be able to deliver germplasm

123
Biodivers Conserv (2013) 22:301–324 317

sustainably, and the establishment of more seed multiplication stands of key species in
farmers’ fields and in public lands. However, some indigenous trees of conservation
concern are difficult to cultivate (López-Gómez et al. 2008) and/or are of little priority to
farmers (Barrance et al. 2009). On the last point for example, of the 58 tree species found
in farmland that were reported by 52 smallholders’ to be of some benefit to them in the
survey of Méndez et al. (2007; Table 1), seven species were of international conservation
concern, but only three were mentioned as being of use by more than one farmer, indi-
cating no common view to support conservation efforts. Clearly, new methods including
the use of markets that bring value to a wider range of species may be required if con-
servation efforts are to be successful (Leakey et al. 2007; Jamnadass et al. 2010). In
addition, information sharing of the ‘service’ and subsistence values of trees species should
not be neglected, as these can be considerable and equally or more important than com-
mercial value, depending on circumstances (Faye et al. 2011).
In the last decade, a participatory approach to tree domestication, in which communi-
ties’ traditional knowledge about tree use and management is combined with scientific
advances in germplasm collection, selection and propagation, and with market develop-
ment (Leakey et al. 2005, 2007), has been applied to bring indigenous fruit and nut trees
(IFTs) from the wild into local cultivation in Central Africa with some success in pro-
moting farm diversification and wider impacts on incomes and health (Tchoundjeu et al.
2010; Asaah et al. 2011; Jamnadass et al. 2011). The approach involves independent
selections of locally ‘elite’ germplasm at multiple sites and so overall genetic losses for a
species during domestication are expected to be low (Leakey et al. 2005, 2007). Fur-
thermore, diversity losses would likely be much greater if domestication leading to
increased productivity did not take place, as the alternative would be loss of the tree from
the entire landscape in competition with other, otherwise more productive, agricultural
options. Whatever strategies are applied to cultivate diversity, however, it seems inevitable
that some interspecific diversity will need to be displaced in currently highly species-rich
agricultural landscapes to maintain the Ne of a smaller range of tree species above a critical
level, in response to the death of remnants and to account for the consequences of climate
change for pollination (Dawson et al. 2011).
A possible response to climate change in agroforestry systems that is not available for
natural forests and woodlands is human-facilitated translocation of germplasm to newly
environmentally suitable planting sites (Dawson et al. 2011). Effective planning of such
transfers requires a good understanding of the current distributions of tree species and the
patterns of adaptive genetic variation found within them, knowledge that has often been
lacking (Aitken et al. 2008). Work to address current gaps in knowledge includes the
development of new and detailed vegetation maps that consider climate change impacts
on vegetation types and tree species through ecological niche modelling (see, e.g.,
Lillesø et al. 2011 and van Breugel et al. 2011 for Africa, MAPFORGEN for Latin
America, www.mapforgen.org/). New trials on indigenous trees valuable to smallholders
are beginning to specifically consider climate change-related traits for populations col-
lected from different ecological zones; for example, the SAFRUIT initiative does so for
Sahelian fruits including baobab (Adansonia digitata) and African locust bean (Parkia
biglobosa) (www.safruit.org; Sanou et al. 2007). Information from these trials will
inform subsequent germplasm distribution strategies that allow farmers to obtain maxi-
mum benefits from tree planting in a changing climate and support circa situm con-
servation. Potential problems of provenance mixing, as local wild populations interact
with better-adapted newly introduced cultivated germplasm, need however to be
addressed.

123
318 Biodivers Conserv (2013) 22:301–324

Acknowledgments We contacted over 60 scientists working on agroforestry and related disciplines to ask
for their views on the topics covered in this essay, and to provide relevant case studies. Many responded,
including Manuel Bertomeu, Charles Clement, Sammy Carsan, Delia Catacutan, Steve Cobb, Ric Coe, Steve
Franzel, Dennis Garrity, Anja Gassner, Lars Graudal, Chris Harwood, Verina Ingram, Hannah Jaenicke,
Katja Kehlenbeck, Roger Leakey, Jens-Peter Lillesø, Seline Meijer, Edward Millard, Simon Mngomba,
Mathew Mpanda, Jonathan Muriuki, Lucy Mwaura, Frank Place, Roberto Porro, Jim Roshetko, Cuauhtémoc
Sáenz-Romero, Kate Schreckenberg, Tony Simons, Pal Singh, Marcos Tito, Julio Ugarte, Patrick van
Damme, Meine van Noordwijk and Maarten van Zonneveld. Charles Clement also reviewed an earlier
version of this manuscript. Several of the authors involved in this paper are part of the CGIAR Research
Programme ‘Forests, trees and agroforestry: livelihoods, landscapes and governance’. This Programme aims
to enhance management and use of forests, agroforestry and tree genetic resources across the landscape,
from forests to farms. This Programme involves the Center for International Forestry Research, Bioversity
International, the International Center for Tropical Agriculture, the World Agroforestry Centre, and others.

References

Aerts R, Berecha G, Gijbels P, Hundera K, Van Glabeke S, Vandepitte K, Muys B, Roldan-Ruiz I, Honnay
O (2013) Genetic variation and risks of introgression in the wild Coffea arabica gene pool in south-
western Ethiopian montane rainforests. Evol Appl (in press)
Aitken SN, Yeaman S, Holliday JA, Wang T, Curtis-McLane S (2008) Adaptation, migration or extirpation:
climate change outcomes for tree populations. Evol Appl 1:95–111
Angelsen A, Kaimowitz D (2004) Is agroforestry likely to reduce deforestation? In: Schroth G, Fonseca
GAB, Harvey CA, Gascon C, Vasconcelos HL, Izac A-MN (eds) Agroforestry and biodiversity con-
servation in tropical landscapes. Island Press, Washington, DC, pp 87–106
Asaah EK, Tchoundjeu Z, Leakey RRB, Takousting B, Njong J, Edang I (2011) Trees, agroforestry and
multifunctional agriculture in Cameroon. Int J Agric Sustain 9:110–119
Balmford A, Green R, Phalan B (2012) What conservationists need to know about farming. Proc R Soc B
279:2714–2724
Barrance A, Schreckenberg K, Gordon J (2009) Conservation through use: lessons from the Mesoamerican
dry forest. Overseas Development Institute, London
Beukema H, Danielson F, Vincent G, Hardiwinoto S, van Andel J (2007) Plant and bird diversity in rubber
agroforests in the lowlands of Sumatra, Indonesia. Agrofor Syst 70:217–242
Bhagwat SA, Willis KJ, Birks HJB, Whittaker RJ (2008) Agroforestry: a refuge for tropical biodiversity?
Trends Ecol Evol 23:261–267
Boshier DH (2004) Agroforestry systems: important components in conserving the genetic viability of
native tropical tree species? In: Schroth G, da Fonseca GAB, Harvey CA, Gascon C, Vasconcelos HL,
Izac A-MN (eds) Agroforestry and biodiversity conservation in tropical landscapes. Island Press,
Washington, DC, pp 290–313
Bourdeix R, Johnson V, Baudouin L, Tuia VS, Kete T, Planes S, Lusty C, Weise S (2011) Polymotu: a new
concept of island-based germplasm bank based on an old Polynesian practice. Ogasawara Research
37:33–51
Brockerhoff EG, Jactel H, Parrotta JA, Quine CP, Sayer J (2008) Plantation forests and biodiversity:
oxymoron or opportunity? Biodivers Conserv 17:925–951
Cassano CR, Schroth G, Faria D, Delabie JHC, Bede L (2009) Landscape and farm scale management to
enhance biodiversity conservation in the cocoa producing region of southern Bahia, Brazil. Biodivers
Conserv 18:577–603
CGIAR Consortium (2011) CGIAR Research Program 6. Forests, trees and agroforestry: livelihoods,
landscapes and governance. The Consultative Group on International Agricultural Research (CGIAR),
CGIAR Consortium Office, Montpellier, France
Charlesworth D, Charlesworth B (1987) Inbreeding depression and its evolutionary consequences. Annu
Rev Ecol Syst 18:237–268
Clapp RA (2001) Tree farming and forest conservation in Chile: do replacement forests leave any originals
behind? Soc Nat Resour 14:341–356
Clement CR (1999) 1492 and the loss of Amazonian crop genetic resources. I. The relation between
domestication and human population decline. Econ Bot 53:188–202
Clement CR, Junqueira AB (2010) Between a pristine myth and an impoverished future. Biotropica
42:534–536

123
Biodivers Conserv (2013) 22:301–324 319

Correia M, Diabaté M, Beavogui P, Guilavogui K, Lamanda N, de Foresta H (2010) Conserving forest tree
diversity in Guinée Forestière (Guinea, West Africa): the role of coffee-based agroforests. Biodivers
Conserv 19:1725–1747
Cossalter C, Pye-Smith C (2003) Fast-wood forestry: myths and realities. The Center for International
Forestry Research (CIFOR), Bogor, Indonesia
Danielsen F, Beukema H, Burgess ND, Parish F, Brühl CA, Donald PF, Murdiyarso D, Phalan B, Reijnders
L, Struebig M, Fitzherbert EB (2009) Biofuel plantations on forested lands: double jeopardy for
biodiversity and climate. Conserv Biol 23:348–358
Dawson I (1997) Prunus africana: how agroforestry can help save an endangered medicinal tree. Agrofor
Today 9(2):15–17
Dawson IK, Jamnadass R (2007) Mainstreaming biodiversity around threatened biodiversity hotspots in
Africa, building on the innovative Allanblackia business. Supporting synthesis for a proposal to the
Global Environment Facility. The World Agroforestry Centre (ICRAF), Nairobi
Dawson IK, Hollingsworth PM, Doyle JJ, Kresovich S, Weber JC, Sotelo-Montes C, Pennington TD,
Pennington RT (2008) Origin and genetic conservation of tropical trees in agroforestry systems: a case
study from the Peruvian Amazon. Conserv Genet 9:361–372
Dawson IK, Lengkeek A, Weber JC, Jamnadass R (2009) Managing genetic variation in tropical trees:
linking knowledge with action in agroforestry ecosystems for improved conservation and enhanced
livelihoods. Biodivers Conserv 18:969–986
Dawson IK, Vinceti B, Weber JC, Neufeldt H, Russell J, Lengkeek AG, Kalinganire A, Kindt R, Lillesø
J-PB, Roshetko J, Jamnadass R (2011) Climate change and tree genetic resource management:
maintaining and enhancing the productivity and value of smallholder tropical agroforestry landscapes.
A review. Agrofor Syst 81:67–78
Delplancke M, Alvarez N, Espindola A, Joly H, Benoit L, Brouck E, Arrigo N (2012) Gene flow among wild
and domesticated almond species: insights from chloroplast and nuclear markers. Evol Appl 5:317–329
Dick CW (2001) Genetic rescue of remnant tropical trees by an alien pollinator. Proc R Soc B 268:
2391–2396
Doerr VAJ, Doerr ED, Davies MJ (2010) Does structural connectivity facilitate dispersal of native species in
Australia’s fragmented terrestrial landscapes? Systematic Review No. 44, Collaboration for Envi-
ronmental Evidence, CSIRO, Canberra, Australia
Donald PF (2004) Biodiversity impacts of some agricultural commodity production systems. Conserv Biol
18:17–37
Doran J, Bush D, Page T, Glencross K, Sethy M, Viji I (2012) Variation in growth traits and wood density in
whitewood (Endospermum medullosum): a major timber species in Vanuatu. Int For Review
14:476–485
Duncan RS, Chapman CA (1999) Seed dispersal and potential forest succession in abandoned agriculture
in tropical Africa. Ecol Appl 9:998–1008
Edmands S (2007) Between a rock and a hard place: evaluating the relative risks of inbreeding and
outbreeding for conservation and management. Mol Ecol 16:463–475
Ekadinata A, Vincent G (2011) Rubber agroforests in a changing landscape: analysis of land use/cover
trajectories in Bungo District, Indonesia. Forests, Trees and Livelihoods 20:3–14
Ellstrand NC (2003) Dangerous liaisons? When cultivated plants mate with their wild relatives. John
Hopkins University Press, Baltimore
FAO (2008) Rapid assessment of pollinators’ status: a contribution to the international initiative for the
conservation and sustainable use of pollinators. Global Action on Pollination Services for Sustainable
Agriculture. Food and Agriculture Organization of the United Nations, Rome
FAO (2010a) Global forest resources assessment 2010. FAO Forestry Paper No. 163, Food and Agriculture
Organization of the United Nations, Rome, Italy
FAO (2010b) The second report on the state of the world’s plant genetic resources for food and agriculture.
Food and Agriculture Organization of the United Nations, Rome
Faye MD, Weber JC, Abasse TA, Boureima M, Larwanou M, Bationo AB, Diallo BO, Sigué H, Dakouo
J-M, Samaké O, Sonogo Diaité D (2011) Farmers’ preferences for tree functions and species in the
West African Sahel. Forests, Trees and Livelihoods 20:113–136
Garen EJ, Saltonstall K, Ashton MS, Slusser JL, Mathias S, Hall JS (2011) The tree planting and protecting
culture of cattle ranchers and small-scale agriculturalists in rural Panama: opportunities for refores-
tation and land restoration. For Ecol Manage 261:1684–1695
Garrity DP (2004) Agroforestry and the achievement of the Millennium Development Goals. Agrofor Syst
61:5–17
Giam X, Bradshaw CJA, Tan HTW, Sodhi NS (2010) Future habitat loss and the conservation of plant
biodiversity. Biol Conserv 143:1594–1602

123
320 Biodivers Conserv (2013) 22:301–324

Gilbert-Norton L, Wilson R, Stevens JR, Beard KH (2010) A meta-analytic review of corridor effectiveness.
Conserv Biol 24:660–668
Gilman RT, Fabina NS, Abbott KC, Rafferty NE (2012) Evolution of plant–pollinator mutualisms in
response to climate change. Evol Appl 5:2–16
Groves CR, Game ET, Anderson MG, Cross M, Enquist C, Ferdana Z, Girvetz E, Gondor A, Hall KR,
Higgins J, Marshall R, Popper K, Schill S, Shafer SL (2012) Incorporating climate change into
systematic conservation planning. Biodivers Conserv 21:1651–1671
Guevara S, Laborde J (1993) Monitoring seed dispersal at isolated standing trees in tropical pastures:
consequences for local species availability. Vegetatio 107(108):319–338
Gyau A, Chiatoh M, Franzel S, Asaah E, Donovan J (2012) Determinants of farmers’ tree planting
behaviour in the North West Region of Cameroon: the case of Prunus africana. International Forestry
Review 14:265–274
Harvey CA, Haber WA (2008) Remnant trees and the conservation of biodiversity in Costa Rican pastures.
Agrofor Syst 44:37–68
Harvey CA, Villalobos JAG (2007) Agroforestry systems conserve species-rich but modified assemblages of
tropical birds and bats. Biodivers Conserv 16:2257–2292
Harwood CE, Bell JC, Moran GF (1992) Isozyme studies on genetic variation and the breeding system in
Grevillea robusta. In: Harwood CE (ed) Grevillea robusta in agroforestry and forestry. The World
Agroforestry Centre (ICRAF), Nairobi, pp 165–176
Harwood CE, Moran GF, Bell JC (1997) Genetic differentiation in natural populations of Grevillea robusta.
Aust J Bot 45:669–678
Hayes TM (2006) Parks, people, and forest protection: an institutional assessment of the effectiveness of
protected areas. World Dev 34:2064–2075
Heywood VH, Dulloo ME (2005) In situ conservation of wild plant species: a critical global review of best
practices. IPGRI Technical Bulletin No. 11. Bioversity International, Rome, Italy
Holl KD (1999) Factors limiting tropical rain forest regeneration in abandoned pasture: seed rain, seed
germination, microclimate and soil. Biotropica 31:229–242
Hollingsworth PM, Dawson IK, Goodall-Copestake WP, Richardson JE, Weber JC, Sotelo-Montes C,
Pennington RT (2005) Do farmers reduce genetic diversity when they domesticate tropical trees? A
case study from Amazonia. Mol Ecol 14:497–501
Homma AKO (1996) Modernisation and technological dualism in the extractive economy in Amazonia. In:
Ruiz-Perez M, Arnold JEM (eds) Current issues in non-timber forest product research. The Center for
International Forestry Research (CIFOR), Bogor, pp 59–81
Jamnadass R, Hanson J, Poole J, Hanotte O, Simons TJ, Dawson IK (2005) High differentiation among
populations of the woody legume Sesbania sesban in sub-Saharan Africa: implications for conserva-
tion and cultivation during germplasm introduction into agroforestry systems. For Ecol Manage
210:225–238
Jamnadass R, Dawson IK, Anegbeh P, Asaah E, Atangana A, Cordeiro N, Hendrickx H, Henneh S, Kadu
CAC, Kattah C, Misbah M, Muchugi A, Munjuga M, Mwaura L, Ndangalasi HJ, Njau CS, Nyame SK,
Ofori D, Peprah T, Russell J, Rutatina F, Sawe C, Schmidt L, Tchoundjeu Z, Simons T (2010)
Allanblackia, a new tree crop in Africa for the global food industry: market development, smallholder
cultivation and biodiversity management. Forests, Trees and Livelihoods 19:251–268
Jamnadass RH, Dawson IK, Franzel S, Leakey RRB, Mithöfer D, Akinnifesi FK, Tchoundjeu Z (2011)
Improving livelihoods and nutrition in sub-Saharan Africa through the promotion of indigenous and
exotic fruit production in smallholders’ agroforestry systems: a review. International Forest Review
13:338–354
Janzen DH (1986) Blurry catastrophes. Oikos 47:1–2
Jha S, Dick CW (2010) Native bees mediate long-distance pollen dispersal in a shade coffee landscape
mosaic. Proc Natl Acad Sci USA 107:13760–13764
Kehlenbeck K, Kindt R, Sinclair FL, Simons AJ, Jamnadass R (2011) Exotic tree species displace indig-
enous ones on farms at intermediate altitudes around Mount Kenya. Agrofor Syst 83:133–147
Kelly BA, Hardy OJ, Bouvet J-M (2004) Temporal and spatial genetic structure in Vitellaria paradoxa (shea
tree) in an agroforestry system in southern Mali. Mol Ecol 13:1231–1240
Kindt R (2002) Methodology for tree species diversification planning in African agroecosystems. PhD
Thesis, University of Gent, Gent, Belgium
Kindt R, Van Damme P, Simons AJ, Beeckman H (2006) Planning tree species diversification in Kenya
based on differences in tree species composition between farms. I. Analysis of tree uses. Agrofor Syst
67:215–228
Kirschenmann FL (2007) Potential for a new generation of biodiversity in agroecosystems of the future.
Agronomy Journal 99:373–376

123
Biodivers Conserv (2013) 22:301–324 321

Kozlowski G, Gibbs D, Huan F, Frey D, Gratzfeld J (2012) Conservation of threatened relict trees through
living ex situ collections: lessons from the global survey of the genus Zelkova (Ulmaceae). Biodivers
Conserv 21:671–685
Kramer AT, Ison JL, Ashley MV, Howe HF (2008) The paradox of forest fragmentation genetics. Conserv
Biol 22:878–885
Kull CA, Tassin J (2012) Australian acacias: useful and (sometimes) weedy. Biol Invasions 14:2229–2233
Labouisse J, Bellachew B, Kotecha S, Bertrand B (2008) Current status of coffee (Coffea arabica L.) genetic
resources in Ethiopia: implications for conservation. Genet Resour Crop Evol 55:1079–1093
Laikre L, Allendorf FD, Aroner LC, Baker CS, Gregovich DP, Hansen MH, Jackson JA, Kendall KC,
Mckelvey K, Neel MC, Olivieri I, Ryman N, Schwartz MK, Bull RS, Jeffrey B, Stetz JB, Tallmon DA,
Taylor BL, Vojta CD, Waller DM, Waples RS (2010) Neglect of genetic diversity in implementation of
the convention of biological diversity. Conserv Biol 24:86–88
Lambert J, Srivastava J, Vietmeyer N (1997) Medicinal plants: rescuing a global heritage. World Bank
Technical Paper 355. The World Bank, Washington DC
Leakey RRB (2010) Agroforestry: a delivery mechanism for multi-functional agriculture. In: Kellimore LR
(ed) Handbook on agroforestry: management practices and environmental impact. Environmental
Science, Engineering and Technology Series. Nova Science Publishers, Hauppauge, pp 461–471
Leakey RRB, Tchoundjeu Z, Schreckenberg K, Shackleton SE, Shackleton CM (2005) Agroforestry tree
products (AFTPs): targeting poverty reduction and enhanced livelihoods. Int J Agric Sustain 3:1–23
Leakey RRB, Tchoundjeu Z, Schreckenberg K, Simons AJ, Shackleton S, Mander M, Wynberg R,
Shackleton C, Sullivan C (2007) Trees and markets for agroforestry tree products: targeting poverty
reduction and enhanced livelihoods. In: Garrity D, Okono A, Parrott M, Parrott S (eds) World agro-
forestry into the future. The World Agroforestry Centre (ICRAF), Nairobi, pp 11–22
Ledig FT (1992) Human impacts on genetic diversity on forest ecosystems. Oikos 63:87–108
Lengkeek AG, Jaenicke H, Dawson IK (2005a) Genetic bottlenecks in agroforestry systems: results of tree
nursery surveys in East Africa. Agrofor Syst 63:149–155
Lengkeek AG, Kindt R, van der Maesen LJG, Simons AJ, van Oijen DCC (2005b) Tree density and
germplasm source in agroforestry ecosystems in Meru, Mount Kenya. Genet Resour Crop Evol
52:709–721
Lengkeek AG, Muchugi Mwangi A, Agufa CAC, Ahenda JO, Dawson IK (2006) Comparing genetic
diversity in agroforestry systems with natural forest: a case study of the important timber tree Vitex
fischeri in central Kenya. Agrofor Syst 67:293–300
Lillesø JB, van Breugel P, Kindt R, Bingham M, Demissew S, Dudley C, Friis I, Gachathi F, Kalema J,
Mbago F, Minani V, Moshi H, Mulumba J, Namaganda M, Ndangalasi H, Ruffo C, Jamnadass R,
Graudal LOV (2011) VECEA: potential natural vegetation of Eastern Africa (Ethiopia, Kenya,
Malawi, Rwanda, Tanzania, Uganda and Zambia). Volume 1: the atlas. Forest & Landscape Working
Paper No. 61. Forest & Landscape, University of Copenhagen, Copenhagen, Denmark
López-Gómez AM, Williams-Linera G, Manson RH (2008) Tree species diversity and vegetation structure
in shade coffee farms in Veracruz, Mexico. Agric Ecosyst Environ 124:160–172
Low T (2012a) Australian acacias: weeds or useful trees? Biol Invasions 14:2217–2227
Low T (2012b) In denial about dangerous aid. Biol Invasions 14:2235–2236
Lowe AJ, Boshier D, Ward M, Bacles CFE, Navarro C (2005) Genetic resource impacts of habitat loss and
degradation; reconciling empirical evidence and predicted theory for neotropical trees. Heredity
95:255–273
Malcolm JR, Markham A, Neilson RP, Garaci M (2002) Estimated migration rates under scenarios of global
climate change. J Biogeogr 29:835–849
Malhi Y, Aragao LEOC, Galbraith D, Huntingford C, Fisher R, Zelazowski P, Sitch S, McSweeney C, Meir
P (2009) Exploring the likelihood and mechanism of a climate change-induced dieback of the Amazon
rainforest. Proc Natl Acad Sci USA 106:20610–20615
Marjokorpi A, Ruokolainen K (2003) The role of traditional forest gardens in the conservation of tree
species in West Kalimantan, Indonesia. Biodivers Conserv 12:799–822
Marshall D, Schreckenberg K, Newton AC (eds) (2006) Commercialization of non-timber forest products:
factors influencing success. Lessons learned from Mexico and Bolivia and policy implications for
decision-makers. UNEP World Conservation Monitoring Centre, Cambridge
Maunder M, Culham A, Alden B, Zizka G, Orliac C, Lobin W, Bordeu A, Ramirez JM, Glissmann-Gough S
(2000) Conservation of the toromiro tree: case study in the management of a plant extinct in the wild.
Conserv Biol 14:1341–1350
McNeely JA (2004) Nature vs. nurture: managing relationships between forests, agroforestry and wild
biodiversity. Agrofor Syst 61:155–165

123
322 Biodivers Conserv (2013) 22:301–324

Méndez VE, Gliessman SR, Gilbert GS (2007) Tree biodiversity in farmer cooperatives of a shade coffee
landscape in western El Salvador. Agric Ecosyst Environ 119:145–159
Miller AJ, Gross BL (2011) From forest to field: perennial fruit crop domestication. Am J Bot 98:1389–1414
Miller AJ, Schaal BA (2005) Domestication of a Mesoamerican cultivated fruit tree, Spondias purpurea.
Proc Natl Acad Sci USA 102:12801–12806
Miller AJ, Schaal BA (2006) Domestication and the distribution of genetic variation in wild and cultivated
populations of the Mesoamerican fruit tree Spondias purpurea L. (Anacardiaceae). Mol Ecol
15:1467–1480
Mohan Jain S, Priyadarshan PM (eds) (2009) Breeding plantation tree crops. Tropical species. Springer,
New York
Murniati, Garrity DP, Gintings AN (2001) The contribution of agroforestry systems to reducing farmers’
dependence on the resources of adjacent national parks: a case study from Sumatra, Indonesia. Agrofor
Syst 52:171–184
Nair PKR, Kumar BM, Nair VD (2009) Agroforestry as a strategy for carbon sequestration. J Plant Nutr Soil
Sci 172:10–23
Newton AC (2008) Conservation of tree species through sustainable use: how can it be achieved in practice?
Oryx 42:195–205
Oldfield SF (2009) Botanic gardens and the conservation of tree species. Trends Plant Sci 14:581–583
Oldfield S, Lusty C, MacKinven A (1998) The world list of threatened trees. World Conservation Press,
Cambridge
Otegbeye GO (1991) Age trends in the genetic-control of stem diameter of Eucalyptus tereticornis and the
implication for selection. Silvae Genetica 40:85–87
Paquette A, Messier C (2010) The role of plantations in managing the world’s forests in the Anthropocene.
Front Ecol Environ 8:27–34
Peprah T, Ofori DA, Siaw DEKA, Addo-Danso SD, Cobbinah JR, Simons AJ, Jamnadass R (2009)
Reproductive biology and characterization of Allanblackia parviflora A. Chev. in Ghana. Genet Resour
Crop Evol 56:1037–1044
Peres CA, Gardner TA, Barlow J, Zuanon J, Michalski F, Lees AC, Vieira ICG, Moreira FMS, Feeley KJ
(2010) Biodiversity conservation in human-modified Amazonian forest landscapes. Biol Conserv
143:2314–2327
Petit RJ, Hampe A (2006) Some evolutionary consequences of being a tree. Annu Rev Ecol Evol Syst
37:187–214
Philpott SM, Bichier P, Rice RA, Greenberg R (2008) Biodiversity conservation, yield, and alternative
products in coffee agroecosystems in Sumatra, Indonesia. Biodivers Conserv 17:1805–1820
Place F, Ajayi OC, Masters E (2011) Tree-based and other land management technologies for landscape
restoration and livelihood in Africa. In: Dewees P, Place F, Scherr SJ, Buss C (principal authors).
Investing in trees and landscape restoration in Africa: what, where, and how. Program on Forests
(PROFOR), Washington DC, pp 17–44
Potts BM, Barbour RC, Hingston AB (2001) The risk of genetic pollution from farm forestry using eucalypt
species and hybrids. Rural Industries Research and Development Corporation, Kingston
Ricketts TH, Daily GC, Ehrlich PR, Michener CD (2004) Economic value of tropical forest to coffee
production. Proc Natl Acad Sci USA 101:12579–12582
Rolim SG, Chiarello AG (2004) Slow death of Atlantic forest trees in cocoa agroforestry in southeastern
Brazil. Biodivers Conserv 13:2679–2694
Ruf FO (2011) The myth of complex cocoa agroforests: the case of Ghana. Human Ecology 39:373–388
Sambuichi RHR, Haridasan M (2007) Recovery of species richness and conservation of native Atlantic
forest trees in the cacao plantations of southern Bahia in Brazil. Biodivers Conserv 16:3681–3701
Sanou H, Korbo A, Tougani A, Rabiou A, Kambou S, Ouedraogo M, Diallo BO, Parkouda C, Ræbild A,
Jensen JS (2007) Protocol for establishment of trials with baobab and tamarind within the SAFRUIT
project. Working Paper No. 21-2007. Forest and Landscape Denmark, Hørsholm, Denmark
Scherr SJ, McNeely JA (2008) Biodiversity conservation and agricultural sustainability: towards a new
paradigm of ‘ecoagriculture’ landscapes. Philos Trans R Soc B 363:477–494
Schippmann U, Leaman DJ, Cunningham AB,(2002) The impact of cultivation and gathering of medicinal
plants on biodiversity: global trends and issues. In: Biodiversity and the ecosystem approach in
agriculture, forestry and fisheries. Satellite event on the occasion of the Ninth Regular Session of the
Commission on Genetic Resources for Food and Agriculture. Rome, 12–13 October 2002. Inter-
Departmental Working Group on Biological Diversity for Food and Agriculture, Rome, Italy
Schmidt L (2007) Tropical forest seed. Springer, Heidelberg
Schoen DJ, Brown AHD (2001) The conservation of wild plant species in seed banks. Bioscience
51:960–966

123
Biodivers Conserv (2013) 22:301–324 323

Schroth G, da Mota MSS (2004) The role of agroforestry in biodiversity conservation in the tropics: a
synthesis of evidence. Paper presented at the International Ecoagriculture Conference and Practitio-
ners’ Fair, held in Nairobi, Kenya. The World Agroforestry Centre (ICRAF), Nairobi, September–
October 2004
Schroth G, Faria D, Araujo M, Bede L, Van Bael SA, Cassano CR, Oliveira LC, Delabie JHC (2011)
Conservation in tropical landscape mosaics: the case of the cacao landscape of southern Bahia, Brazil.
Biodivers Conserv 20:1635–1654
Sheil D, Basuki I, German L, Kuyper TW, Limberg G, Puri RK, Sellato B, van Noordwijk M, Wollenberg E
(2012) Do anthropogenic dark earths occur in the interior of Borneo? Some initial observations from
East Kalimantan. Forests 3:207–229
Simons AJ, Dawson IK, Duguma B, Tchoundjeu Z (1998) Passing problems: prostate and Prunus africana.
Herbalgram 43:49–53
Sonwa DJ, Nkongmeneck BA, Weise SF, Tchatat M, Adesina AA, Janssens MJJ (2007) Diversity of plants
in cocoa agroforests in the humid forest zone of Southern Cameroon. Biodivers Conserv 16:2385–2400
Steffan-Dewenter I, Kessler M, Barkmann J, Bos MM, Buchori D, Erasmi S, Faust H, Gerold G, Glenk K,
Gradstein SR, Guhardja E, Harteveld M, Hertel D, Hohn P, Kappas M, Kohler S, Leuschner C,
Maertens M, Marggraf R, Migge-Kleian S, Mogea J, Pitopang R, Schaefer M, Schwarze S, Sporn SG,
Steingrebe A, Tjitrosoedirdjo SS, Tjitrosoemito S, Twele A, Weber R, Woltmann L, Zeller M,
Tscharntke T (2007) Tradeoffs between income, biodiversity, and ecosystem functioning during
tropical rainforest conversion and agroforestry intensification. Proc Natl Acad Sci USA
104:4973–4978
Strandby-Andersen U, Prado Cordova JP, Nielsen UB, Smith-Olsen C, Nielsen C, Sørensen M, Kollmann J
(2008) Conservation through utilization: a case study of the vulnerable Abies guatemalensis in Gua-
temala. Oryx 42:206–213
Tchoundjeu Z, Degrande A, Leakey RRB, Nimino G, Kemajou E, Asaah E, Facheux C, Mbile P, Mbosso C,
Sado T, Tsobeng A (2010) Impacts of participatory tree domestication on farmer livelihoods in West
and Central Africa. Forests, Trees and Livelihoods 19:217–234
Tscharntke T, Sekercioglu CH, Dietsch TV, Sodhi NS, Hoehn P, Tylianakis JM (2008) Landscape con-
straints on functional diversity of birds and insects in tropical agroecosystems. Ecology 89:944–951
Tscharntke T, Clough Y, Wanger TC, Jackson L, Motzke I, Perfecto I, Vandermeer J, Whitbread A (2012)
Global food security, biodiversity conservation and the future of agricultural intensification. Biol
Conserv 151:53–59
van Breugel P, Kindt R, Lillesø JB, Bingham M, Demissew S, Dudley C, Friis I, Gachathi F, Kalema J,
Mbago F, Minani V, Moshi H, Mulumba J, Namaganda M, Ndangalasi H, Ruffo C, Védaste M,
Jamnadass R, Graudal LOV (2011) VECEA: potential natural vegetation of Eastern Africa (Ethiopia,
Kenya, Malawi, Rwanda, Tanzania, Uganda and Zambia). Volume 7: projected distributions of
potential natural vegetation types and two important agroforestry species (Prunus africana and
Warburgia ugandensis) for six possible future climates. Forest & Landscape Working Paper No. 69.
Forest & Landscape, University of Copenhagen, Copenhagen, Denmark
van Leeuwen J, Lleras Pérez E, Clement CR (2005) Field genebanks may impede instead of promote crop
development: lessons of failed genebanks of ‘‘promising’’ Brazilian palms. Agrociencia, Montevideo
9:61–66
Varghese M, Kamalakannan R, Harwood CE, Lindgren D, McDonald MW (2009) Changes in growth
performance and fecundity of Eucalyptus camaldulensis and E. tereticornis during domestication in
southern India. Tree Genetics and Genomes 5:629–640
Waldron A, Justicia R, Smith L, Sanchez M (2012) Conservation through chocolate: a win-win for biodi-
versity and farmers in Ecuador’s lowland tropics. Conservation Letters 5:213–221
Ward M, Dick CW, Gribel R, Lowe AJ (2005) To self, or not to self… A review of outcrossing and pollen-
mediated gene flow in neotropical trees. Heredity 95:246–254
Weber JC, Labarta-Chávarri RL, Sotelo-Montes C, Brodie AW, Cromwell E, Schreckenberg K, Simons AJ
(1997) Farmers’ use and management of tree germplasm: case studies from the Peruvian Amazon
Basin. In: Simons AJ, Kindt R, Place F (eds) Policy aspects of tree germplasm demand and supply.
Proceedings of an international workshop. The World Agroforestry Centre (ICRAF), Nairobi,
pp 57–63
Weeks AR, Sgro CM, Young AG, Frankham R, Mitchell NJ, Miller KA, Byrne M, Coates DJ, Eldridge
MDB, Sunnucks P, Breed MF, James EA, Hoffmann AA (2011) Assessing the benefits and risks of
translocations in changing environments: a genetic perspective. Evol Appl 4:709–725
White GM, Boshier DH, Powell W (2002) Increased pollen flow counteracts fragmentation in a tropical dry
forest: an example from Swietenia humilis Zuccarini. Proc Natl Acad Sci USA 99:2038–2042

123
324 Biodivers Conserv (2013) 22:301–324

Wiersum KF (1997) Indigenous exploitation and management of tropical forest resources: an evolutionary
continuum in forest–people interactions. Agric Ecosyst Environ 63:1–16
Wiersum KF, Dold AP, Husselman M, Cocks M (2006) Cultivation of medicinal plants as a tool for
biodiversity conservation and poverty alleviation in the Amatola region, South Africa. In: Bogers RJ,
Craker LE, Lange D (eds) Medicinal and aromatic plants. Springer, Dordrecht, pp 43–57
Williams MC, Wardle GM (2009) Pinus radiata invasion in New South Wales: the extent of spread. Plant
Prot Q 24:146–156
Zomer RJ, Trabucco A, Coe R, Place F (2009) Trees on farm: analysis of global extent and geographical
patterns of agroforestry. Working Paper No. 89. The World Agroforestry Centre, Nairobi, Kenya

123

You might also like