Desktop versionMobile version

Conservation Biology in Sub-Saharan Africa

 | 
John W. Wilson
, 
Richard B. Primack

14. Conservation on Unprotected Lands

Full text

Farm workers tending rice fields near Monrovia, Liberia. Agricultural expansion is a major contributor to habitat loss, but is also a common cause of pollution, invasive species, and at scale, even climate change. There is an urgent need to adopt more sustainable agricultural practices and policies to ensure this sector continues to sustain food security and economic activity, but not at the expense of biodiversity.
Photograph by blk24ga, https://commons.wikimedia.org/​wiki/​File:Liberia, _Africa_-_panoramio_(103).jpg, CC BY 3.0.

Well-managed protected areas are essential tools for securing intact ecosystems and the biodiversity they sustain. However, it will not be sufficient to rely solely on protected areas to preserve biodiversity. Harming ecosystems such as rivers and streams on unprotected lands has repeatedly been shown to decrease biodiversity also within protected areas (Colvin et al., 2011; Woodborne et al., 2012). Furthermore, many species only occur on unprotected lands (Beresford et al. 2011), and some species even fare better outside protected areas (Murgatroyd et al., 2016). Other species need to move out of protected areas to access important seasonal resources: about two-thirds of Kenya’s large animals regularly move from protected areas into unprotected rangelands in search of food and water (Young et al., 2005; Western et al. 2009a). Lastly, restoring damaged areas and maintaining intact ecosystems inside and outside of protected areas provides ecosystem services such as water and air purification. And so, while we must continue to pursue protected areas for the many benefits they do offer, we must not forget about the value of the spaces between protected areas. In this chapter, we will explore how efforts on unprotected lands can complement conservation efforts in protected areas.

14.1 Human-Dominated Landscapes

Traditional peoples are important partners in conservation efforts because protecting their lifestyles also ensures the protection of biodiversity.

In every country on Earth, significant portions of unprotected lands still harbour some of their original biota (Figure 14.1). Consider, for example, remote regions that are considered “wilderness” by governments and the general public. Most of these areas are inhabited by low-density human societies that practice a traditional way of life. With relatively little outside influence from modern technology, these traditional peoples are often dependent on—and thus highly concerned with—the health of their environment. More importantly, traditional peoples have been an integral part of their environments for thousands of years. The present mixture and relative densities of wildlife in these “wildernesses” thus reflect the historical activities (e.g. fishing, hunting, fire management, land clearing, and planting of useful plant crops) of the people living in those areas (Roberts et al., 2017). These activities do not degrade the environment if human population densities remain low and natural resources are harvested sustainably. To regulate these activities, most traditional societies have an established system of rights to natural resources, known as customary laws, which an increasing number of governments recognise (Section 12.2.2). Conservation biologists should follow this example: rather than being considered a threat to the “pristine” environments in which they live, traditional peoples should be seen as important partners in conservation efforts because protecting their lifestyles also ensures the protection of biodiversity (Box 14.1).

Figure 14.1 Vertebrate species richness in a variety of land-use systems in West Africa. As human impacts increase, the average number of forest species generally declines while open-habitat species increase.

After Norris et al., 2010, CC BY 4.0.

Box 14.1 Traditional People and Conservation: Turning the Page
Abraham J. Miller-Rushing1 and John W. Wilson
1Acadia National Park, US National Park Service,
Bar Harbor, ME, USA.
How can we balance the conservation of biodiversity with the rights of traditional peoples? Conservation has a mixed history in this regard (Brockington et al., 2006). In many cases, conservation projects and traditional peoples have supported each other, but there are also many examples where vulnerable peoples have been abused and dispossessed of their ancestral lands. Here we briefly discuss three such examples from different regions of Africa: the “Pygmies” of Central Africa’s forests, the “Bushmen” (also known as San, or First People) of the Kalahari Desert, and the Maasai of East Africa’s savannahs.
Relocating local people from protected areas, either before or after establishment, is a relatively common strategy across the world. Amongst other reasons, these relocations are thought of as a strategy to reduce stress on wildlife from hunting and other forms of resource extraction which, in turn, enables park managers to reach their conservation and tourism goals quicker. However, in many cases relocated people are not adequately supported as they transitioned to new lifestyles, leading to increased poverty, declining health, and loss of identity. When vulnerable traditional peoples are relocated, humanity also loses a rich cultural heritage, including local knowledge about native wildlife and traditional medicines.
The Baka, a “Pygmy” tribe that lives in southern Cameroon, have greatly suffered from poorly executed conservation activities. This traditional hunter-gatherer community often identify themselves as “forest peoples” for their strong ties to the forests of Central Africa, which defines their history, culture, and livelihoods. Their ties are now under strain; over the past few decades, several hunting concessions and protected areas were established on ancestral Baka land without proper consultation. In addition to losing their ancestral land, there have also been multiple instances where law enforcement patrols arrested, abused, tortured, and even killed Baka community members that were suspected of illegal hunting (FPP, 2016). Regardless of the back story, it is inherently unjust for hunting to be legal for rich foreigners on ancestral lands where traditional peoples are now prohibited from doing so. This is all even truer when the people who forgo their traditional activities in the name of conservation do not receive compensation for their losses in return.
The second example involves the “Bushmen” (also known as San, or First People). For tens of thousands of years, the Bushmen have lived a nomadic or semi-nomadic hunting and gathering lifestyle in the Kalahari Desert of Southern Africa. Over the last few centuries however, the Bushmen have struggled to maintain their lifestyles as larger African tribes, and later European colonists have staked claims on their land. In 1961, the government of Botswana established the Central Kalahari Game Reserve to protect wildlife and dedicate a place for the Bushmen to practice their traditional lifestyles. However, the government also drilled wells, and established a school and health post which unintentionally encouraged the Bushmen to adopt a more sedentary lifestyle in which they grew crops and raised livestock in the reserve. In 1985, the government of Botswana decided that the Bushmen’s lifestyles were no longer compatible with the goals of the reserve. After a process that most observers, including the UN (Anaya, 2010), considered inadequate and unethical, between 1997 and 2001 the Bushmen were banned from hunting within the reserve, and forced to relocate to settlements outside its boundaries.
In 2006, the High Court of Botswana ruled that the relocations were illegal, and that the government had to allow the Bushmen to return to the reserve. The government, however, allowed only a very limited number of people to return, prohibited traditional hunting, and took other actions that failed to meet their human rights obligations (Shaheed, 2016). The government defended these actions by saying that they were protecting biodiversity, and that the hunting pressure by Bushmen amounted to poaching of protected wildlife. To outside observers, it seemed that keeping them away from a lucrative diamond mine, established in the reserve in 2014, was the stronger motivation (Haines, 2016).
The Maasai are also very familiar with land conflict in the name of conservation. They occupy an area of 160,000 km2 of seasonal lands in the Great Rift Valley in southern Kenya and northern Tanzania. Here they live a semi-nomadic, pastoralist lifestyle, and rely almost entirely on livestock for food and income. In the early 1900s, British colonists began taking lands away from the Maasai to create ranches and later also protected areas. This included the iconic Maasai Mara National Reserve and Serengeti National Park on either side of the Kenya-Tanzania border. Together these parks protect one of the last great mammal migrations in the world, attracting tourists from far and wide.
Maasai pastoralists who live adjacent to these and other protected areas regularly come in conflict with wildlife, particularly lions, which pose a threat to the herders and their livestock. The Maasai have a tradition of hunting lions, particularly after livestock predation, but also as a traditional rite of passage to manhood. Killing lions is now illegal in Kenya as the region’s lion population is dwindling from the impacts of habitat loss and too many lion hunters. As an alternative to retaliatory killings, the Kenyan government financially compensates herders for predation losses; however, the government has not always been consistent in this compensation (Goldman et al., 2010). At times, corrupt officials have also exploited the Maasai by taking or misusing their share of tourism revenues (although corruption seems to be declining and compensation and revenue sharing have improved in recent years). Also, in Tanzania, tens of thousands of Maasai were evicted from their ancestral land in the Serengeti wildlife corridor in 2013. And yet, instead of strict protection, the land in question was earmarked for a foreign-owned hunting concession (Smith, 2014). (The Tanzanian government has since reversed this land grab and fired those behind it.)
Conservation conflicts with traditional peoples are unnecessary and avoidable. Research—done by conservation biologists—shows that traditional peoples have a relatively minor impact on the environment, especially when compared to that of commercial hunters and exploited migrant labourers who have little incentive to live sustainability (Thibault and Blaney, 2003; Poulsen et al., 2009; Fa et al., 2016). Conservation biologists and traditional peoples thus face similar concerns: habitat loss and fragmentation, human encroachment, commercial hunting and other capitalist ventures, and armed militias. Such is the case even in 2019, in southern Cameroon, where the clearing of the Meyomessala forests for Chinese-owned rubber plantations not only threatens many threatened and iconic species, but also the livelihoods of the local Baka community (Sixtus, 2008). As such, working together will benefit both conservation biologists and traditional peoples.
Fortunately, there are several models illustrating the compatibility of conservation and traditional lifestyles. One example comes from northern Kenya, where the community-run Northern Rangeland Trust sustainably manages 44,000 km2 of land. Here, core areas are set aside for wildlife, while buffer areas are subjected to a rotational livestock grazing framework. This model benefits both livestock and wildlife by preventing overgrazing, increasing habitat heterogeneity, and maximising overall biodiversity, thereby allowing 480,000 people (from 15 ethnic groups) living in the area to enjoy peace, prosperity, and healthy ecosystems. Livestock and wildlife populations—which show limited competition under the right conditions (Kartzinel et al., 2015)—are also thriving, and at times they may even benefit one another (Odadi et al., 2011). A major accomplishment here is that the population size of Africa’s rarest antelope, the hirola, doubled in only three and a half years. The area now boasts a growing tourism industry, which boosts employment, and provides revenue that is spent on health, education, and other societal needs.
Some challenges remain, most notably human-wildlife conflict, and overcoming a history of mistrust. Droughts and political instability have also contributed to tensions between conservationists and traditional peoples in some areas. Even so, conservation biologists have an obligation to follow best practices and ensure that human rights are not violated through conservation activities (Borrini-Feyerabend and Hill, 2015). There is also a need to abide by international laws and to recognise the rights of people who want to live a traditional lifestyle; this includes respecting their right to free, prior, and informed consent (FPIC, see UN-REDD, 2013) before mutually-beneficial conservation actions are implemented. With more African countries adopting a western form of land tenure that enables private land ownership, empowering traditional peoples by helping them to obtain legal title (the right to land ownership that is recognised by the government) can also help establish trust and locally managed protected areas where resources are harvested sustainably (Rai and Bawa, 2013). Recent studies have shown that deforestation rates decline on such legally designated traditional lands, particularly in places undergoing rapid land-use changes (Nolte et al., 2013). Gaining trust and working with traditional peoples is not always easy. But only under those circumstances can we have any hope of establishing a sustainable conservation model that will weather the test of the times.

Communal lands in East Africa that are dedicated to pastoralism provide a good example illustrating the compatibility between traditional peoples and conservation efforts (McGahey et al., 2007). In contrast to modern livestock farming systems that maintain livestock in a restricted area, pastoralism involves regularly herding livestock to new areas in search of fresh pasture and water. Pastoralists sometimes also use fire to enhance land productivity, which, together with intense but short-term grazing, maintains native grassland and savannah ecosystems by preventing encroachment of woody plants (Section 10.2.1). By keeping their grazing lands suitable for livestock, pastoralists also maintain those areas in a state that is suitable for native biodiversity and the natural resources they need for survival. In fact, levels of biodiversity on well-managed pastoral land may rival (Msuha et al., 2012) or even exceed (ILRI, 2006) that of adjacent protected areas where such activities are excluded.

Biodiversity conservation and human welfare operate hand-in-hand on unprotected lands, but with a fragile balance that requires continued maintenance.

1However, when the rules governing pastoral systems break down due to agricultural developments and fences that impede pastoralist movements, the resultant breakdown of traditional grazing systems lead to overgrazing, negatively impacting people, livestock, and wildlife (Western et al., 2009b; Groom and Western, 2013). These are the kinds of threats facing the hirola (Beatragus hunteri, CR), also known as the Hunter’s antelope, which shares much of its range with Somali pastoralists in north-eastern Kenya. With fire suppression, elephant extirpation, and a breakdown of traditional grazing systems, trees are encroaching on this facultative grazer’s last remaining strongholds; the loss of grasslands is also harm cattle production in the region (Ali et al., 2017). On pastoralist lands, biodiversity conservation and human livelihoods truly operate hand in hand, but with a fragile balance that requires continued maintenance.

2People who live in rural areas and sell natural resources that they extract from healthy ecosystems also play an important role in conservation by engaging in sustainable natural resource management. Consider all the unprotected estuaries and marine areas that support commercial fisheries for a moment. When fisheries are managed in a sustainable way, not only does this benefit the people that depend on these commercially important species, but other native species can also thrive. Such a win-win outcome was illustrated in Cameroon’s mangrove swamps, where the unsustainable harvesting of mangrove trees for smoking fish was mitigated with fuel-efficient stoves. These new stoves reduced both the amount of time and firewood needed for smoking fish, benefiting the mangroves as well as the fishers’ profit margins (Feka et al., 2009). Fuel-efficient stoves form an important role in reducing firewood harvesting pressure in Cameroon and many other African countries.

3People living in urban centres can also contribute to conservation efforts by raising environmental awareness among fellow citizens, participating in activism, lobbying, and fundraising activities, and generating knowledge through citizen science projects (see Box 15.3). They can also help reduce the multiple pressures that their cities exert on the surrounding environment. Among the most exciting recent developments have been the development and installation of green infrastructure, such as urban forests, green roofs, urban wetlands, permeable sidewalks, urban farms, and rain gardens (Figure 14.2). Not only does green infrastructure save money by reducing energy consumption and pollution clean-up costs, it also reduces overall maintenance (Odefey et al., 2012) and improves overall well-being (Demuzere et al., 2014). Consequently, green infrastructure is increasingly being integrated in urban planning across the world, including North America (EPA, 2018), Europe (Natural England, 2009), Asia (e.g. Kennedy et al., 2016), and in South Africa (Culwick et al., 2016).

Figure 14.2 Green infrastructure enables people living in urban centres to reduce their pressure on natural ecosystems and live more comfortable lives. (Top) A green wall to provide cooling and air purification in Madrid, Spain. Photograph by Jean-Pierre Dalbéra, https://www.flickr.com/​photos/​dalbera/​4657766022, CC BY 2.0. (Bottom) A constructed wetland in Harbin, China, built to provide flood control, a wildlife refuge, and a nature experience.

Photograph by Richard Primack, CC BY 4.0.

Many urbanites are also eager to work with government agencies and conservation NGOs to make their cities more biodiversity friendly by restoring urban waterways and wetlands (Box 14.2), and replanting abandoned industrial sites and other damaged urban areas with native vegetation that can support pollinators, birds, and other wildlife. Such efforts foster neighbourhood pride, create a sense of community, and provide a sense of satisfaction to people who like to be close to nature. These restored areas, and other urban green spaces, can also serve to highlight the links between human well-being and nature, which may make those city dwellers who remain on the side-lines of conservation more receptive to the more challenging aspects of conservation, such as prescribed fire and invasive species management (Gaertner et al., 2016). Establishing and maintaining areas to protect biodiversity where people live and work, termed reconciliation ecology (Rosenzweig, 2003), will increase in importance as Africa’s urban centres continue to expand over the next decades (Seto et al., 2011).

Box 14.2 Importance of Protected Areas in Cities: Insights from the City of Cape Town
Pippin M. L. Anderson
Department of Environmental and Geographical Science,
University of Cape Town, South Africa.
pippin.anderson@uct.ac.za
With urbanisation rates of 3 % per year, Sub-Saharan Africa’s sprawling cities are predicted to increase in area by seven times over coming decades (Anderson et al., 2014). Much of this growth will be associated with weak decentralised governance and limited resources for environmental management. Just under half of all Africans live below the poverty line, and African city-dwellers tend to make greater use of natural resources than citizens of other continents (Anderson et al., 2014). Meanwhile, biodiversity and conservation concerns tend to be eclipsed by, or be independent of, other pressures such as poverty, unemployment, and access to food, water, and housing. Despite the challenges of promoting biodiversity conservation under these conditions, sprawling cities will benefit greatly from conserving some green space within their boundaries.
The City of Cape Town in South Africa has exceptional biodiversity; the region around the city hosts some 9,000 plant species on just 90,000 km2 of land. But as urban areas spread, natural areas are being developed to meet housing and economic needs. Elsewhere, native ecosystems are being suppressed by agricultural transformation, invasive species, annual floods and coastal erosion are intensifying, and altered fire regimes are hindering natural processes. These alterations to the natural landscape have significant economic consequences due to eroded ecosystem services (O’Farrell et al., 2012) conservatively valued at US $150–450 million per annum (de Wit et al., 2009). Similarly, urbanisation also reduces human well-being through loss of education and recreation opportunities afforded by green spaces (Goodness and Anderson, 2014).
The Edith Stephens Nature Reserve, a small (0.39 km2) protected area in the City of Cape Town, illustrates the important role of green space in an urban environment (Figure 14. A). Despite the small size and isolated nature of the reserve, the site provides several benefits to the surrounding community. Originally established to protect a wetland that holds the threatened Cape quillwort (Isoetes capensis, EN), the reserve offers nature-watching opportunities to the public via a bird hide overlooking the wetland, and a quiet neutral space for socialising that is not linked to gang or political territories. The reserve also has an environmental education centre that regularly offers children’s workshops and holiday programmes, as well as a teacher-training programme. Even so, the site faces considerable urban pressures, including heavy foot traffic across the reserve, nutrient run-off from the adjacent agricultural land and urban area, and informal activities, such as brick cleaning on the property’s boundary.

Figure 14. A The Edith Stephens Wetland, South Africa, pictured here with Table Mountain in the background, is surrounded by urban settlement, making it an important reserve for the local communities as it provides green space and ecosystem services. The reserve’s success lies in the fact that it is managed in close collaboration with the adjacent communities.

Photograph by Pippin Anderson, CC BY 4.0.

The reserve was included in the Cape Flats Nature project initiated in 2002, which sought to link urban nature conservation with social justice in the city’s most economically marginal areas (Katzschner, 2013). The idea was to establish conservation practices that integrate ecological sustainability with community empowerment and social well-being. Even though the Cape Flats Nature project ended in 2010, the principles of the project live on through ongoing efforts by the reserve’s small staff and local volunteers recruited from adjacent communities and government public works programmes. The reserve has developed good relations with its neighbours, and the staff claim that they enjoy lower incidences of crime than many other protected areas in the city. The Edith Stephens Nature Reserve exemplifies the kind of hybrid conservation practice that is required in an urban African setting where social and biodiversity requirements need to be balanced, especially in small areas with significant development pressures.

14.1.1 The impact of agriculture

Habitat loss from agricultural expansion is arguably the biggest current challenge to biodiversity conservation in Africa (Balmford et al., 2012; Laurance et al. 2013; Maxwell et al., 2016). At the root of this problem is the need to supply food and other resources to a growing human population. Exacerbating the situation, much of Sub-Saharan Africa’s arable land has already been degraded to such a degree that it cannot sustain viable food production anymore (Drechsel et al., 2001). Most of these losses are not due to natural factors, but to poor land management practices, such as overgrazing, continual ploughing of fields, and heavy use of fertilisers. These practices release nitrogen, carbon, and oxygen into the atmosphere and compromise the soil’s ability to hold water, leading to erosion, soil salinisation, desertification, and even climate change (Vågen et al., 2005). This not only lead to collapsing ecosystem services, but also increased competition for space as even more land must be converted for agriculture, and to accommodate people and their activities. Such land conflicts are only going to become worse with climate change (Zabel et al., 2014).

4In light of the seemingly irreconcilable conflict between agriculture and conservation, some conservation biologists have suggested that the only way in which we can secure a future for biodiversity is through a land sparing approach, in which agricultural investments are focussed on intensifying practices on land already dedicated to farming and no more. One of the main drawbacks of such a high-yield approach is that the impact of intensive agricultural practices degrades natural ecosystems even far from the immediate area, for example through nutrient and pesticide pollution (Section 7.1). For that reason, others support a land sharing approach that promotes biodiversity-friendly agricultural practices, even if that means agricultural lands continue to expand. One of the main drawbacks of this land-sharing approach is that it still alters ecosystem composition, which would threaten species that need large territories and habitats, in addition to leading to more human-wildlife conflict (Section 14.4).

5While the land-sparing versus land-sharing frameworks make for good intellectual debate, the reality is that both are undesirable scenarios when carried out as opposite extremes. There is no denying that agriculture is important—it’s the primary source of livelihood for millions of Africans, and critical for food security. But because there is a finite amount of land available for food production, we have no choice but to develop methods that will allow greater yields on existing agricultural lands without depleting the soil or damaging more ecosystems. In other words, we need to adopt a hybrid approach where some lands are dedicated to large protected areas where human activities are restricted, some lands are dedicated to wildlife-friendly agro-ecosystems, and some lands are used for intensive food production (Fischer et al., 2014; Law et al., 2017). Much of Africa has employed a similar structure historically and even today, where large areas that are relatively untouched or regenerating are interspersed with low-intensity traditional agricultural systems that continue to support a range of native species (Şekercioğlu, 2012). Traditional farming systems offer many strategies showing how natural ecosystem services can be used to improve yields, including the use of biocontrol and crop diversification to keep pests and diseases at bay, and planting of nitrogen-fixing legumes to improve soil fertility. This is in stark contrast to intensive modern agricultural practices dedicated to single crop specialisation; these impoverished ecosystems cannot maintain themselves but rather rely on continuous use of fertilisers and pesticides to remain productive.

Biodiversity-friendly farming practices can also produce economic benefits; and so, many government programmes are now promoting sustainable agricultural intensification.

6With the increased realisation that biodiversity-friendly farming practices can also produce economic benefits, many government programmes have begun to promote and subsidise the adoption of sustainable agricultural intensification (http://www.fao.org/​ag/​ca/​AfricaTrainingManual.html; see also Pretty et al., 2011; Garnett et al., 2013). Also known as conservation agriculture, this farming approach blends traditional agricultural practices with improved and locally adapted crops, as well as integrated crop and pest management strategies (Figure 14.3) to boost yields on existing farmland while creating cost and labour savings; it may even reduce the amount of land under cultivation (Stevenson et al., 2013). Some of these strategies include minimal tilling, crop rotation, intercropping, and terracing (to prevent agricultural runoff which, in turn, prevents erosion). Soil nutrient levels are enhanced through fertiliser microdosing, and by planting legumes, encouraging decomposition by termites, and using crop residues as mulch before composting it directly into the soil. Crop yields are further improved by maintaining windbreaks such as riparian buffer zones, which have the added benefit of enhancing the diversity of seed dispersers, biocontrol agents, and pollinators. In one study, fertiliser microdosing increased sorghum and millet yields and, thus, incomes for 25,000 smallholder farmers in Mali, Burkina Faso, and Niger by 50–130 % (Tabo et al., 2011). Another region-wide study found that conservation agricultural techniques could increase average crop yields by nearly 400 kg per hectare (Corbeels et al., 2014). It is important to remember that conservation agriculture is a departure from both traditional and intensive monocrop farming techniques. Adequately training farmers in best practices and new techniques is therefore crucial to programme success (Gatare et al., 2013).

Figure 14.3 A vegetable farm maintained by a local community living adjacent to Gorongosa National Park, Mozambique. The farm is maintained following several conservation agriculture principles. For example, to conserve water, the ground is covered in mulch obtained from a nearby grassy field, while a drip irrigation system was installed by park staff and a local women’s association. To further strengthen local development and collaboration, part of the crop is also sold at the tourist restaurant inside the park.

Photograph by Iñaki Abella Gutiérrez/Bio+, CC BY. 4.0.

To take advantage of the multiple benefits to be gained from integrated, biodiversity-friendly farming techniques, entire industries have started adopting such practices. Among the most prominent are cacao and coffee, where many growers now produce their crops under native shade trees (Box 14.3). While Africa has had a long history of shade-grown cacao and coffee production, recent decades have seen many farmers transitioning towards intensive farming in full sun, which allows for easier mechanisation. But crops grown in full sun are generally also of lower quality and more susceptible to pest outbreaks (Kellerman et al., 2008; Bisseleua et al., 2009; Tscharntke et al., 2011). In contrast, shade-grown cacao and coffee benefits both the farmers and biodiversity: a study from Ethiopia found that shade coffee farms had over double the number of bird species in comparison with nearby forest sites (Buechley et al., 2015). Shade farming could also be a strategy for farmers trying to cope with increasing temperatures due to climate change (Blaser et al., 2018). Considering the large global markets for cacao and coffee, reverting to traditional growing methods here would have a large positive impact on the environment simply due to the economy of scale.

Box 14.3 Preserving Biodiversity Through Shaded Agroforestry
Hervé D. Bisseleua
World Agroforestry Centre (ICRAF),
Nairobi, Kenya.
hbissel@gmail.com
Chocolate is one of the most universal treats in the world, but could your sweet tooth be increasing biodiversity loss? The chocolate tree (Theobroma cacao) is traditionally grown in areas with dense and diverse canopies of shade trees, home to an abundant variety of plants and animals (Figure 14. B). The chocolate industry is strongly dependent on small-scale agriculture, but also highly vulnerable to pest and disease outbreaks, and climate change. These production challenges combined with increasing global demand for chocolate has increased economic and social pressures to achieve higher yields within a shorter timeframe. Higher yields could be achieved through reduced shade tree management and increased use of chemical pesticides and fertiliser. But these techniques lead to deforestation, biodiversity loss, and loss of ecosystem functioning. Higher yield techniques in the short term are also not sustainable over the long term: work in Cameroon and elsewhere showed that the promotion of high-yielding hybrid cacao varieties under direct full sun have contributed to more frequent outbreaks of pests and diseases (Kellerman et al., 2008; Bisseleua et al., 2009; Tscharntke et al., 2011).
To achieve more sustainable cacao production, and to benefit from ecosystem services, such as enhanced biological control of pests and diseases, and increased soil fertility, West Africa’s cacao farmers are now gradually returning to agroforestry practices that embrace increased shade tree diversity. Farmers adopting these techniques are already reaping benefits. For instance, in Ghana and Cameroon, cacao yields from shaded cacao agroforestry systems are 12–23 % higher compared to full sun systems (Bisseleua et al., 2009; Asare and Raebild, 2016). In eastern Côte d’Ivoire, the use of leguminous trees as shade in rehabilitated cacao plantations is also reported to increase the survival rate and yield of cacao trees (Smith Dumont et al., 2014). Cacao grown in shade may produce for 60–100 years, whereas production may only last for 20 years without shade (Obiri et al., 2007). In addition to environmental services, diversified shade trees may also provide additional income opportunities, such as timber and firewood production, medicine, local spices, and fruit, from native shade trees such as the njangsang tree (Ricinodendron heudelotii) and bush mango (Irvingia gabonensis) (Smith Dumont et al., 2014). Importantly, in all these multi-strata systems, a higher density and diversity of shade trees means higher

Figure 14. B (Top) Cacao grown as a monoculture, without shade trees, in Ghana. Biodiversity is greatly reduced and the cacao trees are more susceptible to pests and diseases in this framework. Photograph by Phillip Allman, CC BY 4.0. (Bottom) Coffee grown under a diverse canopy of trees in Ethiopia. Like cacao, coffee cultivated under shade provides a forest structure in which birds, insects, and other wildlife can flourish.

Photograph by Evan Buechley, CC BY 4.0.

densities and diversity of pollinators and biological pest control agents such as ants and social wasps, which in turn increase cacao yields even more (Bisseleua et al., 2017).
In conclusion, better land management practices, such as allowing a diversity of shade trees to grow among the cacao crop, increases both biodiversity and revenue for farmers. Tropical agroforestry is thus a promising approach to reconcile biodiversity conservation and economic development. Educating farmers on shaded agroforestry systems and creating complimentary economic incentives and policies would help farmers adapt to better management practices faster, ultimately allowing agroforestry systems to contribute more to biodiversity conservation. A guilt-free sweet tooth, indeed!

Sustainable agricultural practices are much needed in the conservation portfolio of Africa, where farmers have been slow to adopt conservation agriculture. Two main challenges are noted. First, spreading ideas and innovations across different agricultural sectors has been challenging, and much work remains to implement them. Second, due to incredible ecosystem diversity across the region, specific practices are not equally suitable everywhere; there is thus need for more research into flexible practices that can be modified to meet local growing conditions (Giller et al., 2009). Strengthening coordination of agricultural research and cooperation at local and regional scales (Gonthier et al., 2014), as well as better land-use allocation (Law et al., 2015) may solve some food production challenges. Adapting biodiversity-friendly certification schemes to consider local dynamics might also encourage more African farmers to adopt biodiversity-friendly techniques (Gove et al. 2008; Buechley et al., 2015). Not only would biodiversity benefit, but these practices would enable farmers across Africa to receive higher prices for their crops and recover a large portion of the land lost to land degradation each year without the need for more land conversion.

14.1.2 The impact of logging, mining, and other extractive industries

As with intensive agriculture, high-impact resource extraction industries have not traditionally been compatible with conservation needs. These include mining, oil and gas extraction, dredging, quarrying, and logging, which have often been associated with complete ecosystem destruction. While it is easy to criticise these industries for their impact on nature, it is important for conservationists to remember that we are all dependent on those industries in some way or other, even to perform our conservation activities. Rather than criticise, it is more productive to partner with and influence with industries to contribute to conservation efforts.

7There are many examples illustrating how partnerships between conservation biologists and extractive industries can benefit conservation. One of the best examples comes from the timber industry, which has the potential to greatly increase forest conservation opportunities (Clark et al., 2009). Traditionally known for leaving unsightly clear-cuts behind them, research has shown that improved logging techniques facilitate quicker ecosystem recovery after harvesting, which in turn also benefits biodiversity. For example, while the response of wildlife to logging differs depending on the harvesting method and forest type (Ofori-Boateng et al., 2013), primates (Stokes et al., 2010; Morgan et al., 2018), amphibians (Ofori-Boateng et al. 2013), and birds (Şekercioğlu, 2002) can all tolerate responsible-done light-touch logging techniques (but see also Bicknell et al., 2013; Gatti et al., 2015).

Conservationists are also dependent on extractive industries. Rather than criticise, it is more productive to partner with and influence these industries to contribute to conservation efforts.

8Guided by this research, some sectors of the timber industry have been keen to adopt more sustainable logging techniques (Figure 14.4) that focus on reducing damage to the soil, stream banks, and remaining trees, while removing just the largest trees. This approach reduces soil disturbance, erosion, waste, and carbon emissions. The Forest Stewardship Council (FSC) and other similar organisations are setting certification standards for sustainable logging, which enable certified logging operations to sell their products at higher prices on world markets. In addition to impact reduction within logged areas, certification schemes typically also require timber companies to avoid logging high conservation value forests, which is a good strategy for protecting ecosystem services and biodiversity in general. Some agroforestry companies also allow local people to cultivate rare medicinal and aromatic plants in the shaded areas on their concessions which reduces harvesting pressure on wild populations (Rao et al., 2004). Lastly, a key element for wildlife management in logged forests is to stop hunters, fishers, trappers, and plant collectors from entering the impacted area after timber harvest by closing unused logging roads. (For more discussion on the impact of the logging road on biodiversity, see Laurance et al., 2014 and Benítez-López et al., 2017.) (For an example in fisheries, see Box 7.2)

9Many African mining companies have also become active partners in conservation. These partnerships include contributions like providing funding for conservation activities, participating in biodiversity offset programmes (Section 10.3.3), and subsidising conservation agriculture efforts. South Africa offers several examples illustrating how extractive industries can develop productive conservation partnerships. Two South African diamond trading companies, De Beers Group and E. Oppenheimer and Son, converted 2,500 km2 of their properties earmarked for diamond mining and exploration into protected areas used for ecotourism and environmental research. Also, in South Africa, the petrochemical company SASOL supports a wide range of environmental programmes, including sponsoring natural history field guides, anti-poaching programs, and threatened species recovery projects.

10Despite these and other examples of progress, many challenges associated with extractive industries remain unsolved. For example, pollution from these industries continues to threaten Africa’s environment, and many extractive companies remain unfriendly towards conservation activities. Many of these challenges stem from efforts to cut costs by ignoring legal and regulatory requirements, particularly when it involves stakeholders from outside Africa who prioritise short-term profits over long-term sustainability. Tardy monitoring and enforcement of regulatory requirements (sometimes driven by corruption) also remain as obstacles (Linder and Palkovitz, 2016). The best solutions for overcoming these challenges involve continued diplomacy to establish new (and strengthen existing) partnerships with such industries. Conservation biologists also need to educate industry workers and the general public to be on the lookout for environmental violations and inadequate law enforcement, which needs to be reported and addressed before more damage is done. Above all, it is important to remind legislators and other members of society that environmental damage can harm, sometimes irreversibly, our own ability to have fulfilling lives.

Figure 14.4 Reduced impact logging techniques facilitate quicker ecosystem recovery. In this example from Mozambique, foresters took only the largest trees, and left some scattered logging slash (i.e. cut branches) to provide shelter for wildlife and to promote natural seed germination. Logged areas are also surrounded by stands of intact forest to promote wildlife and seed dispersal.

Photograph by Johnny Wilson, CC BY 4.0.

14.2 Smart Development Outside Conservation Areas

Infrastructure development poses a significant and escalating challenge to biodiversity conservation efforts. Dams and fences impede wildlife dispersal and migrations (Section 5.1.1), power distribution lines and high-rise buildings pose a collision hazard to birds and bats (Rushworth et al., 2014; Frick et al., 2017), and city expansions compete with biodiversity for space. Expanding road networks are particularly harmful because roads open new areas for deforestation, urban sprawl, agricultural expansion, and unsustainable hunting (Laurence et al., 2006; Benítez-López et al., 2017). A recent review found that 75 % of Sub-Saharan Africa’s development corridors—large-scale infrastructure developments meant to stimulate economic growth—would cut through sparsely-populated and low-quality agricultural, range, and forest lands (Laurance et al., 2015; Sloan et al., 2016). These developments not only threaten the wildlife living within those areas (Benítez-López et al., 2017), but also the carbon-sequestering potential of large swathes of tropical forests (Laurance et al., 2015). Most alarming, the corridors’ planners have shown very little regard for existing biodiversity conservation efforts, given that the proposed transportation network cuts directly through 408 existing protected areas, which includes 69 national parks, biosphere reserves, World Heritage Sites, and Ramsar wetlands (Sloan et al., 2016). In contrast, only five of the 33 planned and active development corridors would cross areas of low conservation priority and with promising agricultural potential (Laurance et al., 2015).

Improving existing infrastructure in disturbed and populated areas makes more economic sense than creating new developments in wilderness areas.

11Conservation versus development is not a zero-sum game. Rather, biodiversity conservation improves our own well-being by enabling us to obtain the necessary resources to support our livelihoods and our industries’ profit margins (Chapter 4). One way to maintain these benefits while also promoting economic development (Section 15.1) is to focus on improving existing infrastructure in disturbed and populated areas, rather than creating new developments that bisect marginal lands, protected areas, and wilderness areas. Developing marginal lands and wildernesses seldom makes sense, not only because these areas are sparsely populated, but also because many are low-nutrient environments that would never support sustainable agriculture (Balmford et al., 2001; Laurance et al., 2015).

12When new developments are necessary, there are usually opportunities to balance diverging interests. For example, to offset the large land footprint of renewable energy, new wind farms and transmission cables could be directed to already degraded land. In many cases, the compromise might even contribute more to socio-economic developmental goals than the original plans. This was well illustrated in Tanzania, where a proposed road development would have disrupted the famous Mara-Serengeti migration route for large mammals, with potentially dire consequences to the area’s ecotourism industry (Dobson et al., 2010; Holdo et al., 2011). To avoid such an impact, scientists used computer models to identify an alternative route that would not only minimise disturbance, but also achieve greater socio-economic development (Hopcraft et al., 2015). Studies, such as these, have provided important foundations for similar work to mitigate the impact of fences on wildlife (Durant et al., 2015), and by making minor adjustments to shipping lanes to reduce collisions between whales and ocean-faring vessels (Silber et al., 2012).

13There are also opportunities to make existing infrastructure more wildlife friendly. Of interest is the maintenance of connectivity despite the presences of potential barriers such as fences and roads. For example, strategically placed fence-gaps and exclusionary fences, as well as tunnels placed under fences can be used to facilitate continued dispersal of selected species in fenced areas (Dupuis-Desormeaux et al., 2018). Similarly, warning signs (Figure 14.5), overpasses (e.g. Ford et al., 2009) and underpasses (e.g. Dell’Amore 2012) along paved roads can keep motorists safe from collisions with large animals. One study from Canada found that strategically placed wildlife crossings could reduce vehicle collisions involving large mammals by 96 % (Ford et al., 2009), also reducing the chance of human injuries and damage to vehicles (Huijser et al., 2009). While this field of research is still relatively new, much headway has been made in making wildlife crossings cost-effective (https://arc-solutions.org) and determining their optimal placement (Bastille-Rousseau et al. 2018).

Figure 14.5 While road signs across Africa warn motorists of dangerous wildlife, the same strategy can be used to protect threatened species. Examples include signs warning motorists about (A) African dogs in Mozambique; (B) threatened frogs in Cape Town, South Africa; (C) threatened dung beetles in South Africa; and (D) rare turtles in South Africa.

Photograph C by Vince Smith, https://www.flickr.com/​photos/​vsmithuk/​2890596997, CC BY 2.0; otherwise by Johnny Wilson, CC BY 4.0.

14.3 Linking Conservation to Socio-Economic Development

Long-term project viability is critically linked to purposeful economic development. Therefore, conservationists are increasingly looking for ways to link conservation to sustainable development, particularly in areas that are impoverished. Integrated conservation and development projects (ICDP) are one of the most popular mechanisms by which this could be accomplished. ICDPs combine conservation activities and local customs with aspects of economic development, including poverty reduction, job creation, health care, and food security. A major goal of ICDPs is for local people to become involved in conservation efforts and have access to opportunities and markets for which sustainable use of natural resources is more valuable than its destructive use. Zambia’s Community Markets for Conservation programme (COMACO) illustrates on how this goal can be achieved (Lewis et al., 2011). Working around the Luangwa Valley’s national parks, COMACO helps food-insecure households and bushmeat hunters to meet their nutritional and income needs through sustainable production of honey, soy, Chama rice, groundnuts, and peanut butter (Figure 14.6). As additional incentive, COMACO connects participants to high-value markets where the villagers’ locally crafted products and sustainably cultivated produce can earn significantly higher prices than locally. Through this project, the area’s average household income has more than quadrupled, over 1,400 bushmeat hunters have adopted more sustainable lifestyles, and over 10,000 km2 of land have been dedicated to community-conserved areas where wildlife populations are now thriving.

Figure 14.6 Many households face a stark choice between poaching/land clearing or food insecurity. As an alternative, COMACO helps rural households in Zambia meet their income and nutritional needs through several sustainable income streams. For example, groundnut farmers are provided opportunities to sell their produce at high-value markets, while groundnut shells (a waste product) are pressed into briquettes that can be used as a renewable fuel source.

Photograph by COMACO, CC BY 4.0.

Community-based natural resource management (CBNRM) represents another approach in which local landowners and community groups can benefit economically from biodiversity and conservation. In previous years, government officials managed biodiversity both inside and outside protected areas through top-down mechanisms with little to no local input. Gaining little economic benefit from the wildlife on their lands, local communities had few incentives to participate in conservation efforts; in some cases, they even became hostile to conservation projects that impeded their activities (Section 13.6.2). To overcome this imbalance, centralised management systems are increasingly transitioning to CBNRM models that involve collaborative management of natural resources on private and communal lands. By empowering local communities and strengthening accountability, government officials and conservation organisations hope that CBNRM projects can simultaneously counterbalance pressures on local wildlife and contribute to economic development in ways that will have long-lasting positive impacts.

14Namibia hosts one of the most ambitious CBNRM projects to date. With seed money from external funders such as the US Agency for International Development (USAID), the Namibian government granted community groups the opportunity to manage the wildlife on their own lands. To obtain these rights, interested community groups needed to form a management committee and determine the boundaries of its land, after which the government designated the group’s land as a “community conservancy”. Participating conservancies then worked with tourist operators— who employed members from the local communities—to provide opportunities for wildlife viewing and hunting (Naidoo et al., 2016), while also allowing tourists to learn about Namibia’s cultural heritage at traditional villages. Revenues from these joint ventures were used to build and maintain even more tourist facilities, and train and pay game guards (also hired from the communal group) who monitor wildlife and human activities on the conservancies. These endeavours have been extremely successful (NACSO, 2015): from the programme’s inception in 1996 to 2014, Namibia’s terrestrial protected area coverage increased from 14 % to 20 %. Wildlife populations also rebounded: for example, Namibia’s elephant population increased from 7,500 to 20,000. Local communities have since reaped the benefits (Störmer et al., 2019). For example, in just 2014, Namibia’s CBNRM projects generated US $6 million in income and provided employment to 5,800 people (NACSO, 2015).

15Unfortunately, maintaining programmes, even successful ones, remains challenging over the long term. Consequently, many previous ICDPs and CBNRM projects have only been partially successful. This includes Zimbabwe’s iconic Communal Areas Management Programme for Indigenous Resources (CAMPFIRE) of the 1990s (Box 14.4), once considered a global model for conservation on unprotected lands. There are many reasons for these projects’ partial successes and failures, including funding limitations, project over-complexity, and political instabilities (Pooley et al., 2014). Although disappointing, these failures have offered valuable lessons that enabled conservation groups to adapt to the challenges of maintaining similar projects over the long term. Today, ICDPs and CBNRM are regarded as worthy of serious consideration, with successful programmes across southern, East, West, and Central Africa (Roe et al., 2009). In addition to providing employment and food security, revenues from ICDPs and CBNRM projects have been used to build schools, clinics, and community centres; improve roads and sanitation; and establish crèches, community gardens, and nurseries (Arntzen et al., 2007; NACSO, 2015). In the end, ICDPs and CBNRM projects will be judged as successful when they can demonstrate that they can both protect wildlife and ensure improved livelihoods over the long term. To achieve these outcomes, a critical component of any ICDP or CBNRM project is the ongoing monitoring of biological, social, and economic factors to determine how effective the programmes are in meeting their goals. Involving local people in these monitoring efforts may increase information sharing and help to determine how aware the people themselves are of the benefits and challenges each project presents (Braschler, 2009).

Box 14.4 Confronting Human-Wildlife Conflict in Zimbabwe
Steven Matema1,2
1African Conservation Trust, Applied Ecology Unit,
Durban, South Africa.
2African Wildlife Economy Institute,
Department of Animal Sciences,
Stellenbosch University, South Africa.
smatemah@gmail.com
In Zimbabwe, several agro-pastoral communities live at the edge of protected areas. They thus come into conflict with wildlife species such as elephants, lions, chacma baboon (Papio ursinus, LC), leopard, spotted hyena (Crocuta crocuta, LC), bushpig (Potamochoerus africana, LC), and common warthog (Phacochoerus africanus, LC) on a regular basis. Zimbabwean law does not provide compensation for crop and livestock losses due to wildlife damage; farmers, thus, develop negative attitudes towards wildlife. Fences and the use of unpalatable buffer crops have not been as successful in mitigating human-wildlife conflict as wildlife conservationists had envisioned (Parker and Osborne, 2006). Instead, lethal control has been the predominant method for managing human-wildlife conflict outside of protected areas, causing a rapid decline in native wildlife populations.
Starting in 1975, the government began to experiment with “people-centred” human-wildlife conflict management strategies (Table 14.1) by adopting the principle that good environmental stewardship is contingent on conferring use and management rights to those directly affected by wildlife depredation. This was the basis for the Communal Areas Management Programme for Indigenous Resources (CAMPFIRE). Under CAMPFIRE, smallholder agro-pastoralists, Rural District Councils (RDC, the land authorities in rural areas), and private safari operators co-manage wildlife outside protected areas and share income from controlled safari hunting and tourism (Murphree, 2009; Taylor, 2009). CAMPFIRE led to a dramatic increase in wildlife populations outside of Zimbabwe’s protected areas: the elephant population increased, and the buffalo population stabilised or declined only slightly outside protected areas (Taylor, 2009). Many of the project’s benefits have also been sustained, despite Zimbabwe’s political volatility over time (Balint and Mashinya, 2008). Yet, in socio-economical terms, CAMPFIRE has largely failed: powerful politicians and local traditional leaders captured benefits, and natural resource governance arrangements have been politicised because political party-affiliated RDC councillors automatically chair local

Table 14.1 Policy and legislative changes for a people-centred approach to wildlife conservation in Zimbabwe, 1975–2005.

Year

Key event

Outcomes for conservation and human-wildlife conflict

1975

Parks and Wildlife Act enacted

Act gives authority to private landowners of white origin to exploit game for profit but leaves out black agro-pastoralists. Wildlife increases on private land. Human-wildlife conflict and negative attitudes toward wildlife by black agro-pastoralists persist.

1978

Wildlife Industries New Development for All (WINDFALL) programme

Culling of meat from parks and distribution to neighbouring communities as a strategy to mitigate human-wildlife conflict improves attitudes towards wildlife. Revenue sent to district councils with no local participation, decision making, or community ownership.

1982

Parks and Wildlife Act amended

The amendment makes provision for authority to be granted to district councils to manage wildlife in rural areas on behalf of the communities.

1984

CAMPFIRE conceived by Department of Parks and Wildlife Management

Target is: collective ownership with defined rights of access to natural resources, appropriate and legitimate institutions, technical and financial assistance.

1989

Authority granted to the first two Rural District Councils

Implementation of CAMPFIRE. Local participation but devolution stops at Rural District Council level.

2005

Direct payment system introduced in CAMPFIRE

Communities receive income due to them from the safari operator directly into a community bank account, bypassing the Rural District Councils (another level of elite capture of income).

Sources: Murphree, 2009; Taylor, 2009.

CAMPFIRE committees following the amendment of the Rural District Councils Act in 2002 (Matema and Andersson, 2015). Ongoing research in the Zambezi Valley also showed that trophy quality has declined since the early 2000s: the horn size of African buffalo (Syncerus caffer, NT) and elephant declined respectively by 42 % (down from 1.35 m to 0.79 m) and 40 % (down from 1.47 m to 0.91 m) between 2006 and 2014 (Matema et al., unpublished litt.). This suggests a decline in the number of adult animals and/or indiscriminate hunting of wildlife, indicating that CAMPFIRE may have also failed to reach its ultimate conservation goals.
Two major lessons were learnt from the CAMPFIRE experience. First, if community-based conservation is to be effective as a human-wildlife conflict mitigation strategy, attention needs to be paid to local and national political dynamics. Second, devolution—the transfer of decision power to local levels—is important. The enactment of Zimbabwe’s Indigenisation and Economic Empowerment Act (2008), which makes provision for rural communities to form Community Share Ownership Trusts to exploit natural resources in their areas, provided a model that CAMPFIRE could have adopted to achieve complete devolution. However, the political elite has used the 2008 Act to demand shares in, or a complete take-over of, wildlife conservancies owned by ranchers of white origin. Communities living next to these conservancies have been excluded in these take-overs with the concomitant escalation of human-human conflict about wildlife (Nyahunzvi, 2014), and negative implications for local tolerance of wildlife species that kill livestock and damage crops. To curb elite capture of income, global compacts are needed, such as the recent ban of imports of wildlife trophies into the USA until there is evidence that local people are equitably sharing revenue from CAMPFIRE. The CAMPFIRE model can work and create greater tolerance for wildlife so long as it buffers local people from income losses. That means compensation in lieu of retaliation on species damaging crops and killing livestock.

14.4 Confronting Human-Wildlife Conflict

As a growing human population continues to encroach on the last remaining wildernesses, wildlife populations are facing increased competition for space and food. Inevitably, as animals are being displaced from degraded ecosystems, they will increasingly meet humans. Some of these interactions will be negative ranging from direct conflict (e.g. injury and even death to one or both participants) and indirect conflict (e.g. transmission of diseases) to opportunity costs (e.g. loss of income due to crop damage and livestock predation). Although human-wildlife conflict is not unique to Africa, Africans are generally very vulnerable due to high levels of poverty and dependence on land, which limits options for conflict mitigation. Managing human-wildlife conflict is thus an important issue to consider in the management of potentially dangerous species, especially near protected area borders.

14.4.1 Dealing with predators

When wildlife impedes human activities, the traditional solution is to either kill the animal or to exclude it from the area with a barrier such as a fence. Killing problem animals can take the form of pro-active lethal control to avoid losses, or retaliatory killings in response to losses. While there is a sense of instant gratification after killing a problem animal, it provides only a temporary solution at best, and may even give rise to a new set of challenges. For example, work on black-backed jackals (Canis mesomelas, LC) showed that killing territorial individuals may cause a breakdown in their local social structure, in turn allowing multiple roaming sub-adult animals to take advantage of the vacant territory (Minnie et al., 2016). Killing apex predators could also give rise mesopredator release, where medium-sized carnivores and omnivores (e.g. jackals and baboons) flourish in the absence of their natural enemies (Brashares et al., 2010). Indiscriminate poisoning and trapping also kills beneficial non-target animals that opportunistically scavenge, such as owls, vultures, and harmless ant-eating mammals (Brown, 2006; Ogada et al., 2015). Thus, while killing problem animals may seem an intuitive solution, it is seldom the best strategy.

Non-lethal control of problem animals involved in human-wildlife conflict may provide more benefits than lethal control.

16Pastoralist communities are particularly vulnerable to predators. Because they are nomadic, pastoralists do not always have access to permanent or sturdy structures to protect their livestock and themselves. Consequently, conservation biologists are spending considerable energy on finding predatory-friendly approaches that offer lasting solutions for pastoralist communities. Among the most successful are schemes that provide compensation payments to pastoralists who forego retaliatory killings following livestock losses (Dickman et al. 2011). In Kenya, for example, compensation schemes reduced retaliatory killings of lions by 73–91 % (Maclennan et al., 2009; Hazzah et al., 2014). Retaliatory killings can be reduced even more when compensation schemes are combined with other strategies; for example, one study that encouraged the use of mobile enclosures for livestock, communal herding, and “lion guardians” (the latter drawing on local knowledge and traditional values to mitigate conflict) saw a drop of 99 % in retaliatory killings (Hazzah et al., 2014).

17Livestock on commercial and smallholder farms are also vulnerable to predation when foraging away from protective structures. Non-lethal options to reduce livestock losses under these circumstances include predator-proof fences using native thorny plants, corralling pregnant females and calves during their vulnerable periods (Schiess-Meier et al., 2007), and setting up visual, chemical, or acoustic repellents in predation hotspots. Eliminating poor livestock husbandry (Woodroffe and Frank, 2005; Gusset et al., 2009; Newsome et al., 2015) and tardy disposal of deceased animals (Humphries et al., 2015) can also avoid situations where predators are attracted to domestic animals in the first place. But perhaps one of the most successful programmes has been the use of livestock guarding animals, which could be dogs (Figure 14.7), donkeys, and other domesticated animals trained to protect livestock. A study from South Africa found that livestock depredation was eliminated on 91 % of farms after the placement of guardian dogs, saving each of the 94 participant farms US $3,189 per year (Rust et al., 2013); Namibian farmers reported equally encouraging results with guardian animals (Marker et al., 2005). While there is an upfront cost involved in obtaining a guardian animal, recent work found that their deployment is generally more efficient and cost-effective than the cost of lethal options (McManus et al., 2015).

Figure 14.7 Placing large guard dogs with livestock is a highly effective, non-lethal strategy to reduce predator attacks. The mere presence of the dog is often enough to keep predators away. Several conservation organisations are now providing trained guard dogs to reduce instances of human-wildlife conflicts involving predators.

Photograph by Cheetah Conservation Fund, https://en.wikipedia.org/​wiki/​File:Kangal_Shepherd_ (livestock-guarding_dog)_and_flock_of_goats_ in_Namibia.jpg, CC BY-SA 3.0.

The collaborations between farmers and conservation biologists to reduce livestock predation have benefited biodiversity conservation as well. Populations of African wild dogs (Lycaon pictus, EN) and lions are rebounding on some unprotected lands (Woodruffe, 2011; Blackburn et al., 2016), while farmers using guardian animals are also more tolerant of some predators on their properties (Rust et al., 2013). Not only do these farmers enjoy seeing native wildlife on their properties; some have even completely switched focus away from livestock to more profitable ecotourism (Sims-Castley et al., 2005) and wildlife ranching endeavours (Lindsey et al., 2013).

14.4.2 Dealing with crop raiders

Non-lethal management of crop-raiding animals is also a high priority. The traditional non-lethal method of dealing with potentially dangerous crop-raiding species (e.g. elephants) involves maintaining electric fences (Kioko et al., 2008), but this method is expensive and requires electricity. To overcome these challenges, conservationists and communities have developed several innovative strategies that may even supplement incomes. One such method is to establish buffer fences made of honey-producing beehives (Scheijen et al., 2019) or chilli plants (Parker and Osborn, 2006; Chang’a et al., 2016); tea plants have also been used successfully to keep crop-raiding gorillas (Gorilla spp.) at bay (Seiler and Robbins, 2016). Using a different approach, conservation biologists in Tanzania developed a harmless, low-cost alarm kit to deter elephants (Bale, 2016). This four-step system involves first shining bright flashing lights at an approaching elephant, followed by loud air horns, then launching a grenade filled with chilli powder, sand, and a loud firecracker, and, as a last resort, launching exploding fireworks toward the approaching elephant.

14.4.3 Concluding thoughts on human-wildlife conflict

One of the most effective mechanisms for dealing with human-wildlife conflict is to develop awareness and opportunities for at-risk people to benefit from potentially harmful animals.

Whether dealing with dangerous animals or crop raiders, one of the most effective mechanisms for dealing with human-wildlife conflict is to develop awareness and opportunities for at-risk people to benefit from potentially harmful animals (Blackburn et al., 2016). Studies in northern Ethiopia found that most people—even those who have been victims of human-wildlife conflict—have positive attitudes towards wildlife and believe that they can co-exist (Eshete et al., 2015). The reason for such positive attitudes is that a substantial portion of the affected people are aware of benefits from ecosystem services, including ecotourism opportunities. Such positive attitudes toward wildlife play a crucial role in the protection of a range of endemic species in this Global Biodiversity Hotspot, including the Walia ibex (Capra walie, EN) and Ethiopian wolf (Canis simensis, EN).

18As discussed earlier, both ICDPs and CBNRM programmes offer opportunities for local people to gain direct benefits from local wildlife, even potentially dangerous species. There are also research opportunities to further human-wildlife conflict mitigation beyond direct benefits to local people. For example, much progress has been made in understanding how lion (Tuqa et al., 2014) and elephant (Granados et al., 2012; Chiyo et al., 2014) behaviours relate to human activities; a logical next step would be to use this information to reduce conflict (e.g. Packer et al., 2005). An increasing number of resources are available to aid these and other efforts. The IUCN Human-Wildlife Conflict Task Force has taken the lead to collate much of this information; their library (http://www.hwctf.org/​resources/​document-library) is sorted by species and topic. They also provide free training manuals (e.g. Parker et al., 2007) and host regular workshops.

14.5 Summary

  1. Many species persist outside protected areas, in areas such as traditional farmland, sustainably logged forests, and unprotected rangelands. These areas can and must play a more important role in ongoing conservation efforts.

  2. Traditional peoples that live on undeveloped land have beliefs that are compatible with biodiversity conservation. There are conservation strategies that can benefit traditional people and protect biodiversity.

  3. Areas intensively used by humans can also contribute to conservation efforts. Biodiversity-friendly techniques are being developed and implemented for the agriculture, logging, and fisheries industries, many which have been adopted. Mines and other extractive industries can participate in biodiversity offset programmes, and partner with conservation organisations to contribute to local biodiversity protection. But there remains a need to monitor the activities of extractive industries to ensure that cost-cutting measures do not lead to biodiversity losses.

  4. Integrated conservation and development projects (ICDPs) and community-based natural resource management (CBNRM) projects link biodiversity conservation with economic development. There is however a need to ensure these approaches remain resilient to challenges that may threaten their long-term success.

  5. Human-wildlife conflict, such as livestock predation and crop raiding, remains a major conservation challenge. Multiple mechanisms have been developed to help victims overcome or mitigate such losses. Some of these mechanisms have even allowed victims to benefit from the animals they previously thought of as nuisances.

14.6 Topics for Discussion

  1. Imagine that the government informs you that a highly threatened species lives on land that you planned to develop. Would you be happy, angry, confused, or proud? What are your options in terms of the planned development? What would be a fair compromise that would protect your rights and interests, the rights of the public, and the well-being of the species?

  2. Imagine your country builds an expensive dam to provide hydroelectricity and water for irrigation. It will take decades to pay back the costs of construction and lost ecosystem services; some of those costs may never be recovered. Who are the winners of such a project, and who are the losers? How are each of these groups (consider both people and wildlife groups) affected? What do you think can be done to make the project more worthwhile?

  3. Do you think that the purchase of “green” (environmentally-responsible) products is an effective way to promote biodiversity conservation? Would you be willing to spend more money for timber, fuelwood, coffee, chocolate, palm oil, and other products that have been produced in a sustainable way, and if so, how much more? How could you determine whether the purchase of such products was really making a difference?

  4. Think of a family (someone you know or heard of) that has been a victim of human-wildlife conflict or contracted a disease while being in nature. What happened? What did the family lose? Was the family compensated for their losses? How and by whom? If you had the opportunity to establish a plan to prevent or mitigate future conflicts, what would you do?

14.7 Suggested Readings

Balmford, A., R. Green, and B. Phalan. 2012. What conservationists need to know about farming. Proceedings of the Royal Society B 279: 2714–24. https://doi.org/​10.1098%2Frspb.2012.0515 Farming is the basis of modern civilisation but can also be damaging to nature.

Cox, R.L., and E.C. Underwood. 2011. The importance of conserving biodiversity outside of protected areas in Mediterranean ecosystems. PLoS ONE 6: e14508. https://doi.org/​10.1371/​journal.pone.0014508 Unprotected lands have the potential to contribute to an overall conservation strategy.

Hassanali, A., H. Herren, Z.R. Khan, et al. 2008. Integrated pest management: The push-pull approach for controlling insect pests and weeds of cereals, and its potential for other agricultural systems including animal husbandry. Philosophical Transactions of the Royal Society of London B 363: 611–21. https://doi.org/​10.1098/​rstb.2007.2173 Benefits of integrated pest management strategies extend beyond pest control and increased crop yields.

Hopcraft, J.G.C., S.A.R. Mduma, M. Borner, et al. 2015. Conservation and economic benefits of a road around the Serengeti. Conservation Biology 29: 932–36. https://doi.org/​10.1111/​cobi.12470 Compromises between conservation and development might contribute more to socio-economic developmental goals than the original plans.

Laurance, W.F., S. Sloan, L. Weng, et al. 2015. Estimating the environmental costs of Africa’s massive “development corridors”. Current Biology 25: 3202–08. https://doi.org/​10.1016/​j. cub.2015.10.046 Il-conceived development wastes resources and harms biodiversity.

Lewis, D., S.D. Bell, J. Fay, et al. 2011. Community Markets for Conservation (COMACO) links biodiversity conservation with sustainable improvements in livelihoods and food production. Proceedings of the National Academy of Sciences 108: 13957–62. https://doi.org/​10.1073/​pnas.1011538108 An example of a programme that links conservation with socio-economic upliftment.

McManus, J.S., A.J. Dickman, D. Gaynor, et al. 2015. Dead or alive? Comparing costs and benefits of lethal and non-lethal human-wildlife conflict mitigation on livestock farms. Oryx 49: 687–95. https://doi.org/​10.1017/​S0030605313001610 Non-lethal methods provide more benefits than lethal methods in controlling predators of livestock.

Morgan, D., R. Mundry, C. Sanz, et al. 2018. African apes coexisting with logging: Comparing chimpanzee (Pan troglodytes troglodytes) and gorilla (Gorilla gorilla gorilla) resource needs and responses to forestry activities. Biological Conservation 218: 277–86. https://doi.org/​10.1016/​j. biocon.2017.10.026 Guidance for sustainable logging aimed at protecting apes.

Bibliography

Pretty, J., C. Toulmin, and S. Williams. 2011. Sustainable intensification in African agriculture. International Journal of Agricultural Sustainability 9: 5–24. https://doi.org/10.3763/ijas.2010.0583 Sustainable agricultural intensification benefits conservation and food security.

Bibliography

Ali, A.H., A.T. Ford, J.S. Evans, et al. 2017. Resource selection and landscape change reveal mechanisms suppressing population recovery for the world’s most endangered antelope. Journal of Applied Ecology 54: 1720–29. https://doi.org/10.1111/1365-2664.12856

Anaya, J. 2010. Report of the Special Rapporteur on the situation of human rights and fundamental freedoms of indigenous people. Addendum. The situation of indigenous peoples in Botswana (New York: UNHRC). https://undocs.org/A/HRC/15/37/Add.2

Anderson, P.M.L., C. Okereke, A. Rudd, et al. 2014. Regional assessment of Africa. In: Urbanization, Biodiversity and Ecosystem Services: Challenges and Opportunities: A Global Assessment, ed. by T. Elmqvist, et al. (New York: Springer). https://doi.org/10.1007/978-94-007-7088-1

Arntzen, J., T. Setlhogile, and J. Barnes. 2007. Rural livelihoods, poverty reduction and food security in Southern Africa: Is CBNRM the answer? (Washington: USAID). http://unpan1.un.org/intradoc/groups/public/documents/cpsi/unpan026980.pdf

Asare, A., and A. Raebild. 2016. Tree diversity and canopy cover in cocoa systems in Ghana. New Forests 47: 287–302. https://doi.org/10.1007/s11056-015-9515-3

Bale, R. 2016. How chili condoms and firecrackers can help save elephants. National Geographic. http://on.natgeo.com/28SSeWj

Balint, P.J., and J. Mashinya. 2008. CAMPFIRE during Zimbabwe’s national crisis: Local impacts and broader implications for community-based wildlife management. Society and Natural Resources 21: 783–96. https://doi.org/10.1080/08941920701681961

Balmford, A., J.L. Moore, T. Brooks, et al. 2001. Conservation conflicts across Africa. Science 291: 2616–19. https://doi.org/10.1126/science.291.5513.2616

Balmford, A., R. Green, and B. Phalan. 2012. What conservationists need to know about farming. Proceedings of the Royal Society B 279: 2714–24. https://doi.org/10.1098/rspb.2012.0515

Bastille-Rousseau, G., J. Wall, I. Douglas-Hamilton et al. 2018. Optimizing the positioning of wildlife crossing structures using GPS telemetry. Journal of Applied Ecology 55: 2055–63. https://doi.org/10.1111/1365-2664.13117

Benítez-López, A., R. Alkemade, A.M. Schipper, et al. 2017. The impact of hunting on tropical mammal and bird populations. Science 356: 180–83. https://doi.org/10.1126/science.aaj1891

Beresford, A.E., G.M. Buchanan, P.F. Donald, et al. 2011. Poor overlap between the distribution of protected areas and globally threatened birds in Africa. Animal Conservation 14: 99–107. https://doi.org/10.1111/j.1469-1795.2010.00398.x

Bicknell, J.E., M.J. Struebig, D.P. Edwards, et al., 2013. Improved timber harvest techniques maintain biodiversity in tropical forests. Current Biology 24: R1119–20. https://doi.org/10.1016/j.cub.2014.10.067

Bisseleua, H.B.D., A.D. Missoup, and S. Vidal. 2009. Biodiversity conservation, ecosystem functioning, and economic incentives under cocoa agroforestry intensification. Conservation Biology 23: 1176–84. https://doi.org/10.1111/j.1523-1739.2009.01220.x

Bisseleua, H.B.D., D. Begoude, H. Tonnang, et al. 2017. Ant-mediated ecosystem services and disservices on marketable yield in cocoa agroforestry systems. Agriculture, Ecosystems and Environment 247: 407–17. https://doi.org/10.1016/j.agee.2017.07.004

Blackburn, S., J.G.C. Hopcraft, J.O. Ogutu, et al. 2016. Human-wildlife conflict, benefit sharing and the survival of lions in pastoralist community-based conservancies. Journal of Applied Ecology 53: 1195–205. https://doi.org/10.1111/1365-2664.12632

Blaser, W.J., J. Oppong, S.P. Hart, et al., 2018. Climate-smart sustainable agriculture in low-to-intermediate shade agroforests. Nature Sustainability 1: 234–39. https://doi.org/10.1038/s41893-018-0062-8

Borrini-Feyerabend, G., and R. Hill. 2015. Governance for the conservation of nature. In: Protected Area Governance and Management, ed. by G.L. Worboys, et al. (Canberra: ANU Press). http://doi.org/10.22459/PAGM.04.2015

Braschler, B. 2009. Successfully implementing a citizen-scientist approach to insect monitoring in a resource-poor country. BioScience 59: 103–04. https://doi.org/10.1525/bio.2009.59.2.2

Brashares J.S., C.W. Epps, and C.J. Stoner. 2010. Ecological and conservation implications of mesopredator release. In: Trophic Cascades, ed. by J. Terborgh and J. Estes (Washington: Island Press).

Brockington, D., J. Igoe, and K. Schmidt-Soltau. 2006. Conservation, human rights, and poverty reduction. Conservation Biology 20: 250–52. https://doi.org/10.1111/j.1523-1739.2006.00335.x

Brown,C.J.2006. HistoricdistributionoflargemammalsintheGreaterFishRiverCanyonComplex, southern Namibia, and recommendations for re-introductions (Windhoek: Namibia Nature Foundation). http://www.the-eis.com/data/literature/Greater%20Fish%20River%20Canyon%20Complex%20Historic%20distribution%20of%20mammals.pdf

Buechley, E.R., Ç. H. Şekercioğlu, A. Atickem, et al. 2015. Importance of Ethiopian shade coffee farms for forest bird conservation. Biological Conservation 188: 50–60. https://doi.org/10.1016/j. biocon.2015.01.011

Chang’a, A., N. de Souza, J. Muya, et al. 2016. Scaling-up the use of chili fences for reducing human-elephant conflict across landscapes in Tanzania. Tropical Conservation Science 9: 921– 30. https://doi.org/10.1177%2F194008291600900220

Chiyo, P. I., J.W. Wilson, E.A. Archie, et al. 2014. The influence of forage, protected areas, and mating prospects on grouping patterns of male elephants. Behavioral Ecology 25: 1494–504. https://doi.org/10.1093/beheco/aru152

Clark, C.J., J.R. Poulsen, R. Malonga, et al. 2009. Logging concessions can extend the conservation estate for Central African tropical forests. Conservation Biology 23: 1281–93. https://doi.org/10.1111/j.1523-1739.2009.01243.x

Colvin, C., A. Burns, K. Schachtschneider, et al. 2011. Coal and water futures in South Africa: The case for protecting headwaters in the Enkangala grasslands (Cape Town: WWF-SA). http://awsassets.wwf.org.za/downloads/wwf_coal_water_report_2011_web.pdf

Corbeels M, R.K. Sakyi, R.F. Kühne, et al. 2014. Meta-analysis of crop responses to conservation agriculture in sub-Saharan Africa. CCAFS Report No. 12 (Copenhagen: CCAFS). https://ccafs.cgiar.org/publications/meta-analysis-crop-responses-conservation-agriculture-sub-saharan-africa

Culwick, C., and K. Bobbins. 2016. A framework for a green infrastructure planning approach in the Gauteng City-Region. GCRO Research Report 04 (Johannnesburg: GCRO). http://www.gcro.ac.za/media/reports/GCRO_Green_Assets_Report_Digital_version__book.pdf de Wit, M., H. van Zyl, D. Crookes, et al. 2009. Investing in Natural Assets: A business case for the environment in the City of Cape Town (Cape Town: City of Cape Town).

Dell’Amore, C. 2012. Pictures: Elephant underpass reuniting Kenya herds. National Geographic. http://on.natgeo.com/Lkubxo

Demuzere, M., K. Orru, O. Heidrich, et al. 2014. Mitigating and adapting to climate change: Multi-functional and multi-scale assessment of green urban infrastructure. Journal of Environmental Management 146: 107–15. https://doi.org/10.1016/j.jenvman.2014.07.025

Dickman, A.J., E.A. Macdonald, and D.W. Macdonald. 2011. A review of financial instruments to pay for predator conservation and encourage human-carnivore coexistence. Proceedings of the National Academy of Sciences 108: 13937–44. https://doi.org/10.1073/pnas.1012972108

Dobson, A.P., M. Borner, A.R.E. Sinclair, et al. 2010. Road will ruin Serengeti. Nature 467: 272– 73. https://doi.org/10.1038/467272a

Drechsel, P., L. Gyiele, D. Kunze, et al. 2001. Population density, soil nutrient depletion, and economic growth in sub-Saharan Africa. Ecological Economics 38: 251–58. https://doi.org/10.1016/S0921-8009 (01) 00167-7

Dupuis-Desormeaux, M., T.N. Kaaria, M. Mwololo, et al. 2018. A ghost fence-gap: Surprising wildlife usage of an obsolete fence crossing. PeerJ 6: e5950. https://dx.doi.org/10.7717%2Fpeerj.5950

Durant, S.M., M.S. Becker, S. Creel, et al. 2015. Developing fencing policies for dryland ecosystems. Journal of Applied Ecology 52: 544–51. https://doi.org/10.1111/1365-2664.12415

Edwards, D.P., S. Sloan, L. Weng, et al. 2014. Mining and the African environment. Conservation Letters 7: 302–11. https://doi.org/10.1111/conl.12076

EPA (Environmental Protection Agency). 2018. Green Infrastructure. https://www.epa.gov/green-infrastructure

Eshete, G., G. Tesfay, H. Bauer, et al. 2015. Community resource uses and Ethiopian wolf conservation in Mount Abune Yosef. Environmental Management 56: 684–94. https://doi.org/10.1007/s00267-015-0529-6

Fa, J.E., J. Olivero, M.A. Farfán, et al. 2016. Differences between Pygmy and non-Pygmy hunting in Congo Basin forests. PLoS ONE 11: e0161703. https://doi.org/10.1371/journal.pone.0161703

Feka, N.Z., G.B. Chuyong, and G.N. Ajonina. 2009. Sustainable utilization of mangroves using improved fish-smoking systems: A management perspective from the Douala-Edea wildlife reserve, Cameroon. Tropical Conservation Science 2: 450–68. https://doi.org/10.1177%2F194008290900200406

Fischer, J., D.J. Abson, V. Butsic, et al. 2014. Land sparing versus land sharing: Moving forward. Conservation Letters 7: 149–57. https://doi.org/10.1111/conl.12084

Ford, A.T., A.P. Clevenger, and A. Bennett. 2009. Comparison of methods of monitoring wildlife crossing-structures on highways. Journal of Wildlife Management 73: 1213–22. https://doi.org/10.2193/2008-387

FPP (Forest Peoples Programme). 2016. FPP Statement on survival international’s complaint against WWF. FPP Press Release. https://www.forestpeoples.org/en/news-article/2016/fpp-statement-survival-internationals-complaint-against-wwf

Frick, W.F., E.F. Baerwald, J.F. Pollock, et al. 2017. Fatalities at wind turbines may threaten population viability of a migratory bat. Biological Conservation 209: 172–77. https://doi.org/10.1016/j.biocon.2017.02.023

Gaertner, M., B.M.H. Larson, U.M. Irlich, et al. 2016. Managing invasive species in cities: A framework from Cape Town, South Africa. Landscape and Urban Planning 151: 1–9. https://doi.org/10.1016/j.landurbplan.2016.03.010

Garnett, T., M. Appleby, A. Balmford, et al. 2013. Sustainable intensification in agriculture: Premises and policies. Science 341: 33–34. http://doi.org/10.1126/science.1234485

Gatere, L., J. Lehmann, S. DeGloria, et al. 2013. One size does not fit all: Conservation farming success in Africa more dependent on management than on location. Agriculture, Ecosystems, and Environment 179: 200–07. https://doi.org/10.1016/j.agee.2013.08.006

Gatti, R.C., S. Castaldi, J.A. Lindsell, et al. 2015. The impact of selective logging and clearcutting on forest structure, tree diversity and above-ground biomass of African tropical forests. Ecological Research 30: 119–32. https://doi.org/10.1007/s11284-014-1217-3

Giller, K.E., E. Witter, M. Corbeels, et al. 2009. Conservation agriculture and smallholder farming in Africa: The heretics’ view. Field Crops Research 114: 23–34. https://doi.org/10.1016/j. fcr.2009.06.017

Goldman, N., J.R. de Pinho, and J. Perry. 2010. Maintaining complex relations with large cats: Maasai and lions in Kenya and Tanzania. Human Dimensions of Wildlife 15: 332–46. https://doi.org/10.1080/10871209.2010.506671

Gonthier, D.J., K.K. Ennis, S. Farinas, et al. 2014. Biodiversity conservation in agriculture requires a multi-scale approach. Proceedings of the Royal Society B 281: 20141358. https://doi.org/10.1098/rspb.2014.1358

Goodness, J., and P.M.L. Anderson. 2014. Local assessment of Cape Town. In: Urbanization, Biodiversity and Ecosystem Services: Challenges and Opportunities: A Global Assessment, ed. by T. Elmqvist, et al. (New York: Springer). https://doi.org/10.1007/978-94-007-7088-1

Gove, A.D., K. Hylander, S. Nemomisa, et al. 2008. Ethiopian coffee cultivation—Implications for bird conservation and environmental certification. Conservation Letters 1: 208–16. https://doi.org/10.1111/j.1755-263X.2008.00033.x

Granados, A., R.B. Weladji, and M.R. Loomis. 2012. Movement and occurrence of two elephant herds in a human-dominated landscape, the Bénoué Wildlife Conservation Area, Cameroon. Tropical Conservation Science 52: 150–62. https://doi.org/10.1177%2F194008291200500205

Groom, R., and D. Western. 2013. Impact of land subdivision and sedentarization on wildlife in Kenya’s southern rangelands. Rangeland Ecology and Management 66: 1–9. https://doi.org/10.2111/REM-D-11-00021.1

Gusset, M., M.J. Swarner, L. Mponwane, et al. 2009. Human-wildlife conflict in northern Botswana: livestock predation by endangered African wild dog Lycaon pictus and other carnivores. Oryx 43: 67–72. https://doi.org/10.1017/S0030605308990475

Haines, G. 2016. Travellers urged to boycott Botswana over its treatment of Bushmen. Telegraph. http://tgr.ph/hid7nl

Hazzah, L., S. Dolrenry, L. Naughton, et al. 2014. Efficacy of two lion conservation programs in Maasailand, Kenya. Conservation Biology 28: 851–60. https://doi.org/10.1111/cobi.12244

Holdo, R.M., J.M. Fryxell, A.R.E. Sinclair, et al. 2011. Predicted impact of barriers to migration on the Serengeti wildebeest population. PLoS ONE 6: e16370. https://doi.org/10.1371/journal. pone.0016370

Hopcraft, J.G.C., S.A.R. Mduma, M. Borner, et al. 2015. Conservation and economic benefits of a road around the Serengeti. Conservation Biology 29: 932–36. https://doi.org/10.1111/cobi.12470

Huijser, M.P., J.W. Duffield, A.P. Clevenger, et al. 2009. Cost-benefit analyses of mitigation measures aimed at reducing collisions with large ungulates in the United States and Canada; a decision support tool. Ecology and Society 14: 15. https://www.ecologyandsociety.org/vol14/iss2/art15

Humphries, B.D., T.R. Hill, and C.T. Downs. 2015. Landowners’ perspectives of black-backed jackals (Canis mesomelas) on farmlands in KwaZulu-Natal, South Africa. African Journal of Ecology 53: 540–49. https://doi.org/10.1111/aje.12247

ILRI (International Livestock Research Institute). 2006. Pastoralism: The surest way out of poverty in East African drylands (Nairobi: ILRI). https://cgspace.cgiar.org/handle/10568/2274

Kartzinel, T.R., P.A. Chen, T.C. Coverdale, et al. 2015. DNA metabarcoding illuminates dietary niche partitioning by African large herbivores. Proceedings of the National Academy of Sciences 112: 8019–24. https://doi.org/10.1073/pnas.1503283112

Katzschner, T. 2013. Cape Flats Nature: Rethinking urban ecologies. In: Contested Ecologies: Dialogues in the South on Nature and Knowledge, ed. by L.J.F. Green (Cape Town: HSRC Press).

Kellermann, J.L., M.D. Johnson, A.M. Stercho, et al. 2008. Ecological and economic services provided by birds on Jamaican Blue Mountain coffee farms. Conservation Biology 22: 1177–85. https://doi.org/10.1111/j.1523-1739.2008.00968.x

Kennedy, C., M. Zhong, and J. Corfee-Morlot. 2016. Infrastructure for China’s Ecologically Balanced Civilization. Engineering 2: 414–25. https://doi.org/10.1016/J.ENG.2016.04.014

Kioko, J., P. Muruthi, P. Omondi, et al. 2008. The performance of electric fences as elephant barriers in Amboseli, Kenya. South African Journal of Wildlife Research 38: 52–58. https://doi.org/10.3957/0379-4369-38.1.52

Laurance, W.F., B.M. Croes, L. Tchignoumba, et al. 2006. Impacts of roads and hunting on central African rainforest mammals. Conservation Biology 20: 1251–61. https://doi.org/10.1111/j.1523-1739.2006.00420.x

Laurance, W.F., J. Sayer, and K.G. Cassman. 2014. Agricultural expansion and its impacts on tropical nature. Trends in Ecology and Evolution 29: 107–16. https://doi.org/10.1016/j. tree.2013.12.001

Laurance, W.F., S. Sloan, L. Weng, et al. 2015. Estimating the environmental costs of Africa’s massive “development corridors”. Current Biology 25: 3202–08. https://doi.org/10.1016/j. cub.2015.10.046

Law, E.A., B.A. Bryan, E. Meijaard, et al. 2017. Mixed policies give more options in multifunctional tropical forest landscapes. Journal of Applied Ecology 54: 51–60. https://doi.org/10.1111/1365-2664.12666

Law, E.A., E. Meijaard, B.A. Bryan, et al. 2015. Better land-use allocation outperforms land sparing and land sharing approaches to conservation in Central Kalimantan, Indonesia. Biological Conservation 186: 276–86. https://doi.org/10.1016/j.biocon.2015.03.004

Lewis, D., S.D. Bell, J. Fay, et al. 2011. Community Markets for Conservation (COMACO) links biodiversity conservation with sustainable improvements in livelihoods and food production. Proceedings of the National Academy of Sciences 108: 13957–62. https://doi.org/10.1073/pnas.1011538108

Linder, J.M., and R.E. Palkovitz. 2016. The threat of industrial oil palm expansion to primates and their habitats. In: Ethnoprimatology, ed. by M. Waller (Cham: Springer). https://doi.org/10.1007/978-3-319-30469-4

Lindsey, P.A., C.P. Havemann, R. Lines, et al. 2013. Determinants of persistence and tolerance of carnivores on Namibian ranches: Implications for conservation on southern African private lands. PloS ONE 8: e52458. https://doi.org/10.1371/journal.pone.0052458

Maclennan S.D., R.J. Groom, D.W. Macdonald, et al. 2009. Evaluation of a compensation scheme to bring about pastoralist tolerance of lions. Biological Conservation 11: 2419–27. https://doi.org/10.1016/j.biocon.2008.12.003

Marker, L.L., A.J. Dickman, and D.W. Macdonald. 2005. Perceived effectiveness of livestock-guarding dogs placed on Namibian farms. Rangeland Ecology and Management 58: 329–36. https://doi.org/10.2111/1551-5028 (2005) 058 [0329: PEOLDP] 2.0. CO; 2

Matema, S., and J.A. Andersson. 2015. Why are lions killing us? Human-wildlife conflict and social discontent in Mbire District, northern Zimbabwe. Journal of Modern African Studies 53: 93–120. https://doi.org/10.1017/S0022278X14000664

Maxwell, S.L., R.A. Fuller, T.M. Brooks, et al. 2016. The ravages of guns, nets and bulldozers. Nature 536: 143–45. https://doi.org/10.1038/536143a

McGahey, D., J. Davies, and E. Barrow. 2008. Pastoralism as conservation in the Horn of Africa: Effective policies for conservation outcomes in the drylands of Eastern Africa. Annals of Arid Zones 46: 353–77.

McManus, J.S., A.J. Dickman, D. Gaynor, et al. 2015. Dead or alive? Comparing costs and benefits of lethal and non-lethal human-wildlife conflict mitigation on livestock farms. Oryx 49: 687–95. https://doi.org/10.1017/S0030605313001610

Minnie, L., A. Gaylard, and G.I.H. Kerley. 2016. Compensatory life-history responses of a mesopredator may undermine carnivore management efforts. Journal of Applied Ecology 53: 379–87. https://doi.org/10.1111/1365-2664.12581

Morgan, D., R. Mundry, C. Sanz, et al. 2018. African apes coexisting with logging: Comparing chimpanzee (Pan troglodytes troglodytes) and gorilla (Gorilla gorilla gorilla) resource needs and responses to forestry activities. Biological Conservation 218: 277–86. https://doi.org/10.1016/j. biocon.2017.10.026

Msuha, M.J., C. Carbone, N. Pettorelli, et al. 2012. Conserving biodiversity in a changing world: Land use change and species richness in northern Tanzania. Biodiversity and Conservation 21: 2747–59. http://doi.org/10.1007/s10531-012-0331-1

Murgatroyd, M., L.G. Underhill, L. Rodrigues, et al. 2016. The influence of agricultural transformation on the breeding performance of a top predator: Verreaux’s Eagles in contrasting land use areas. Condor 118: 238–52. https://doi.org/10.1650/CONDOR-15-142.1

Murphree, M.W. 2009. The strategic pillars of communal natural resource management: Benefit, empowerment and conservation. Biodiversity and Conservation 18: 2551–62. https://doi.org/10.1007/s10531-009-9644-0.

NACSO (Namibian Association of CBNRM Support Organisations). 2015. The state of community conservation in Namibia—A review of communal conservancies, community forests and other CBNRM initiatives (Windhoek: NACSO). http://www.nacso.org.na/sites/default/files/2014-15_SoCC-Report.pdf

Naidoo, R., L.C. Weaver, R.W. Diggle, et al. 2016. Complementary benefits of tourism and hunting to communal conservancies in Namibia. Conservation Biology 30: 628–38. https://doi.org/10.1111/cobi.12643

Natural England. 2009. Natural England’s green infrastructure guidance (NE176) (York: Natural England). http://publications.naturalengland.org.uk/publication/35033

Newsome, T.M., J.A. Dellinger, C.R. Pavey, et al. 2015. The ecological effects of providing resource subsidies to predators. Global Ecology and Biogeography 24: 1–11. https://doi.org/10.1111/geb.12236

Nolte, C., A. Agrawal, K.M. Silvus, et al. 2013. Governance regime and location influence avoided deforestation success of protected areas in the Brazilian Amazon. Proceedings of the National Academy of Sciences 110: 4956–5961. https://doi.org/10.1073/pnas.1214786110

Norris, K., A. Asase, B. Collen, et al. 2010. Biodiversity in a forest-agriculture mosaic—The changing face of West African rainforests. Biological Conservation 143: 2341–50. https://doi.org/10.1016/j.biocon.2009.12.032

Nyahunzvi, D.K. 2014. The resurgence in resource nationalism and private protected areas: Through the lens of Save Valley Conservancy’s indigenisation. Journal of Nature Conservation 22: 343–46. https://doi.org/10.1016/j.jnc.2013.08.003

O’Farrell, P.J., P.M.L. Anderson, D. le Maitre, et al. 2012. Insights and opportunities offered by a rapid ecosystem service assessment in promoting a conservation agenda in an urban biodiversity hotspot. Ecology and Society 17: 8. http://doi.org/10.5751/ES-04886-170327

Obiri, B.D., G.A. Bright, M.A McDonald, et al. 2007. Financial analysis of shaded cocoa in Ghana. Agroforestry Systems 71: 139–49. https://doi.org/10.1007/s10457-007-9058-5

Odadi, W.O., M.K. Karachi, S.A. Abdulrazak, et al. 2011. African wild ungulates compete with or facilitate cattle depending on season. Science 333: 1753–55. https://doi.org/10.1126/science.1208468

Odefey, J., S. Detwiler, K. Rousseau, et al. 2012. Banking on green: A look at how green infrastructure can save municipalities money and provide economic benefits community-wide (Washington: American Rivers and others). https://www.americanrivers.org/conservation-resource/banking-on-green

Ofori-Boateng, C., W. Oduro, A. Hillers, et al. 2013. Differences in the effects of selective logging on amphibian assemblages in three west African forest types. Biotropica 45: 94–101. https://doi.org/10.1111/j.1744-7429.2012.00887.x

Ogada, D., A. Botha, and P. Shaw. 2015. Ivory poachers and poison: Drivers of Africa’s declining vulture populations. Oryx: 1–4. https://doi.org/10.1017/S0030605315001209

Packer, C., D. Ikanda, B. Kissui, et al. 2005. Lion attacks on humans in Tanzania — understanding the timing and distribution of attacks on rural communities will help to prevent them. Nature 436: 927–28. https://doi.org/10.1038/436927a

Parker G.E., F.V. Osborn, R.E. Hoare, et al. 2007. Human-elephant conflict mitigation: A training course for community-based approaches in Africa (Livingstone: Elephant Pepper Development Trust; Nairobi: IUCN/SSCAfESG). https://www.iucn.org/sites/dev/files/import/downloads/heccombaptmen.pdf

Parker, G.E., and F.V. Osborn. 2006. Investigating the potential for chilli Capsicum spp. to reduce human-wildlife conflict in Zimbabwe. Oryx 40: 343–46. https://doi.org/10.1017/S0030605306000822

Pooley, S., J.A. Mendelsohn, and E.J. Milner-Gulland. 2014. Hunting down the chimera of multiple disciplinarity in conservation science. Conservation Biology 28: 22–32. https://doi.org/10.1111/cobi.12183

Poulsen, J.R., C.J. Clark, G. Mavah, et al. 2009. Bushmeat supply and consumption in a tropical logging concession in northern Congo. Conservation Biology 23: 1597–608. https://doi.org/10.1111/j.1523-1739.2009.01251.x

Pretty, J., C. Toulmin, and S. Williams. 2011. Sustainable intensification in African agriculture. International Journal of Agricultural Sustainability 9: 5–24. https://doi.org/10.3763/ijas.2010.0583

Rai, N.D., and K.S. Bawa. 2013. Inserting politics and history in conservation. Conservation Biology 27: 425–28. https://doi.org/10.1111/cobi.12026

Rao, M.R., M.C. Palada, and B.N. Becker. 2004. Medicinal and aromatic plants in agroforestry systems. Afroforestry Systems 61: 107–22. https://doi.org/10.1023/B:AGFO.0000028993.83007.4b

Roberts, P., C.O. Hunt, M. Arroyo-Kalin, et al. 2017. The deep human prehistory of global tropical forests and its relevance for modern conservation. Nature Plants 3: 17093. https:// doi.org/10.1038/nplants.2017.93

Roe, D., F. Nelson, and C. Sandbrook. 2009. Community Management of Natural Resources in Africa: Impacts, Experiences and Future Directions. Natural Resource Issues 18 (London: IIED). http://pubs.iied.org/17503IIED

Rosenzweig, M.L. 2003. Reconciliation ecology and the future of species diversity. Oryx 37: 194– 205. https://doi.org/10.1017/S0030605303000371

Rushworth, I., and S. Krüger. 2014. Wind farms threaten southern Africa’s cliff-nesting vultures. Ostrich 8: 13–23. http://doi.org/10.2989/00306525.2014.913211

Rust, N.A., T.M. Whitehouse-Tedd, and D.C. MacMillan. 2013. Perceived efficacy of livestock-guarding dogs in South Africa: Implications for cheetah conservation. Wildlife Society Bulletin 37: 690–97. https://doi.org/10.1002/wsb.352

Scheijen, C.P.J., S.A. Richards, J. Smit, et al., 2019. Efficacy of beehive fences as barriers to African elephants: A case study in Tanzania. Oryx 52: 92–99. https://doi.org/10.1017/S0030605317001727

Schiess-Meier, M., S. Ramsauer, T. Gabanapelo, et al. 2007. Livestock predation—Insights from problem animal control registers in Botswana. Journal of Wildlife Management 71: 1267–74. https://doi.org/10.2193/2006-177

Seiler, N., and M.M. Robbins. 2015. Factors influencing ranging on community land and crop raiding by mountain gorillas. Animal Conservation 19: 176–88. https://doi.org/10.1111/acv.12232

Şekercioğlu, C.H. 2002. Effects of forestry practices on vegetation structure and bird community of Kibale National Park, Uganda. Biological Conservation 107: 229–40. http://doi.org/10.1016/S0006-3207(02)00097-6

Şekercioğlu, C.H. 2012. Bird functional diversity and ecosystem services in tropical forests, agroforests and agricultural areas. Journal of Ornithology 153: 153–61. https://doi.org/10.1007/s10336-012-0869-4

Seto, K.C., M. Fragkias, B. Günerapl, et al. 2011. A meta-analysis of global urban land expansion. PloS ONE 6: e23777. https://doi.org/10.1371/journal.pone.0023777

Shaheed, F. 2016. Report of the Special Rapporteur in the field of cultural rights on her visit to Botswana (New York: UNHRC). http://daccess-ods.un.org/access.nsf/Get?Open&DS=A/HRC/31/59/Add.1&Lang=E

Silber, G.K., A.S.M. Vanderlaan, A.T. Arceredillo, et al. 2012. The role of the International Maritime Organization in reducing vessel threat to whales: Process, options, action and effectiveness. Marine Policy 36: 1221–33. https://doi.org/10.1016/j.marpol.2012.03.008

Sims-Castley, R., G.I. Kerley, B. Geach, et al. 2005. Socio-economic significance of ecotourism-based private game reserves in South Africa’s Eastern Cape Province. Parks 15: 6–18.

Sixtus, M. 2018. Indigenous communities at risk as Chinese rubber firm uses land. Al Jazeera http://aje.io/83rm4

Sloan, S., B. Bertzky, and W.F. Laurance. 2017. African development corridors intersect key protected areas. African Journal of Ecology 55: 731–37. https://doi.org/10.1111/aje.12377

Smith Dumont, E., G.M. Gnahoua, L. Ohouo, et al. 2014. Farmers in Côte d’Ivoire value integrating tree diversity in cocoa for the provision of ecosystem services. Agroforestry Systems 88: 1047–66. https://doi.org/10.1007/s10457-014-9679-4

Smith, D. 2014. Tanzania accused of backtracking over sale of Masai’s ancestral land. Guardian. https://gu.com/p/43cxc

Stevenson, J.R., N. Villoria, D. Byerlee, et al. 2013. Green Revolution research saved an estimated 18 to 27 million hectares from being brought into agricultural production. Proceedings of the National Academy of Sciences 110: 8363–68. https://doi.org/10.1073/pnas.1208065110

Stokes, E.J., S. Strindberg, P. C. Bakabana, et al. 2010. Monitoring great ape and elephant abundance at large spatial scales: Measuring effectiveness of a conservation landscape. PLoS ONE 4: e10294. https://doi.org/10.1371/journal.pone.0010294

Störmer, N., L.C. Weaver, G. Stuart-Hill, et al. 2019. Investigating the effects of community-based conservation on attitudes towards wildlife in Namibia. Biological Conservation 233: 193–200. https://doi.org/10.1016/j.biocon.2019.02.033

Tabo, R, A. Bationo, B. Amadou, et al. 2011. Fertilizer microdosing and “Warrantage” or inventory credit system to improve food security and farmers’ income in West Africa. In: Innovations as Key to the Green Revolution in Africa, ed. by A. Bationo, et al. (Dordrecht: Springer). https:// doi.org/10.1007/978-90-481-2543-2

Taylor, R. 2009. Community based natural resources management in Zimbabwe: the experience of CAMPFIRE. Biodiversity and Conservation 18: 2563–83. https://doi.org/10.1007/s10531-0099612-8

Thibault, M., and S. Blaney. 2003. The oil industry as an underlying factor in the bushmeat crisis in Central Africa. Conservation Biology 17: 1807–13. https://doi.org/10.1111/j.1523-1739.2003.00159.x

Tscharntke, T., Y. Clough, S.A. Bhagwat, et al. 2011. Multifunctional shade-tree management in tropical agroforestry landscapes—a review. Journal of Applied Ecology 48: 619–29. https://doi.org/10.1111/j.1365-2664.2010.01939.x

Tuqa, J.H., P. Funston, C. Musyoki, et al. 2014. Impact of severe climate variability on lion home range and movement patterns in the Amboseli ecosystem, Kenya. Global Ecology and Conservation 2: 1–10. https://doi.org/10.1016/j.gecco.2014.07.006

UN-REDD. 2013. Guidelines on free, prior, and informed consent (Geneva: UN-REDD). https://theredddesk.org/resources/guidelines-free-prior-and-informed-consent

Vågen, T.-G., R. Lal, and B.R. Singh. 2005. Soil carbon sequestration in sub-Saharan Africa: a review. Land Degradation and Development 16: 53–71. https://doi.org/10.1002/ldr.644

Western, D., R. Groom, and J. Worden. 2009b. The impact of subdivision and sedentarization of pastoral lands on wildlife in an African savanna ecosystem. Biological Conservation 142: 2538–46. https://doi.org/10.1016/j.biocon.2009.05.025

Western, D., S. Russell, and I. Cuthill. 2009a. The status of wildlife in protected areas compared to non-protected areas of Kenya. PLoS One 4: e6140. https://doi.org/10.1371/journal. pone.0006140

Woodborne, S., K.D.A. Huchzermeyer, D. Govender, Det al. 2012. Ecosystem change and the Olifants River crocodile mass mortality events. Ecosphere 3: 1–17. https://doi.org/10.1890/ES12-00170.1

Woodroffe, R. 2011 Demography of a recovering African wild dog (Lycaon pictus) population. Journal of Mammalogy 92: 305–15. https://doi.org/10.1644/10-MAMM-A-157.1

Woodroffe, R., and L.G. Frank. 2005. Lethal control of African lions (Panthera leo): Local and regional population impacts. Animal Conservation 8: 91–98. https://doi.org/10.1017/S1367943004001829

Young, T.P., T.M. Palmer, and M.E. Gadd. 2005. Competition and compensation among cattle, zebras, and elephants in a semi-arid savanna in Laikipia, Kenya. Biological Conservation 112: 251–59. https://doi.org/10.1016/j.biocon.2004.08.007

Zabel F, B. Putzenlechner, and W. Mauser. 2014. Global agricultural land resources—A high resolution suitability evaluation and its perspectives until 2100 under climate change conditions. PLoS ONE 99: e107522. https://doi.org/10.1371/journal.pone.0107522

List of illustrations

URL http://books.openedition.org/obp/docannexe/image/10462/img-1.jpg
File image/jpeg, 92k
Caption Figure 14.1 Vertebrate species richness in a variety of land-use systems in West Africa. As human impacts increase, the average number of forest species generally declines while open-habitat species increase.
URL http://books.openedition.org/obp/docannexe/image/10462/img-2.jpg
File image/jpeg, 33k
Caption Figure 14.2 Green infrastructure enables people living in urban centres to reduce their pressure on natural ecosystems and live more comfortable lives. (Top) A green wall to provide cooling and air purification in Madrid, Spain. Photograph by Jean-Pierre Dalbéra, https://www.flickr.com/​photos/​dalbera/​4657766022, CC BY 2.0. (Bottom) A constructed wetland in Harbin, China, built to provide flood control, a wildlife refuge, and a nature experience.
URL http://books.openedition.org/obp/docannexe/image/10462/img-3.jpg
File image/jpeg, 92k
Caption Figure 14. A The Edith Stephens Wetland, South Africa, pictured here with Table Mountain in the background, is surrounded by urban settlement, making it an important reserve for the local communities as it provides green space and ecosystem services. The reserve’s success lies in the fact that it is managed in close collaboration with the adjacent communities.
URL http://books.openedition.org/obp/docannexe/image/10462/img-4.jpg
File image/jpeg, 44k
Caption Figure 14.3 A vegetable farm maintained by a local community living adjacent to Gorongosa National Park, Mozambique. The farm is maintained following several conservation agriculture principles. For example, to conserve water, the ground is covered in mulch obtained from a nearby grassy field, while a drip irrigation system was installed by park staff and a local women’s association. To further strengthen local development and collaboration, part of the crop is also sold at the tourist restaurant inside the park.
URL http://books.openedition.org/obp/docannexe/image/10462/img-5.jpg
File image/jpeg, 66k
Caption Figure 14. B (Top) Cacao grown as a monoculture, without shade trees, in Ghana. Biodiversity is greatly reduced and the cacao trees are more susceptible to pests and diseases in this framework. Photograph by Phillip Allman, CC BY 4.0. (Bottom) Coffee grown under a diverse canopy of trees in Ethiopia. Like cacao, coffee cultivated under shade provides a forest structure in which birds, insects, and other wildlife can flourish.
URL http://books.openedition.org/obp/docannexe/image/10462/img-6.jpg
File image/jpeg, 211k
Caption Figure 14.4 Reduced impact logging techniques facilitate quicker ecosystem recovery. In this example from Mozambique, foresters took only the largest trees, and left some scattered logging slash (i.e. cut branches) to provide shelter for wildlife and to promote natural seed germination. Logged areas are also surrounded by stands of intact forest to promote wildlife and seed dispersal.
URL http://books.openedition.org/obp/docannexe/image/10462/img-7.jpg
File image/jpeg, 61k
Caption Figure 14.5 While road signs across Africa warn motorists of dangerous wildlife, the same strategy can be used to protect threatened species. Examples include signs warning motorists about (A) African dogs in Mozambique; (B) threatened frogs in Cape Town, South Africa; (C) threatened dung beetles in South Africa; and (D) rare turtles in South Africa.
URL http://books.openedition.org/obp/docannexe/image/10462/img-8.jpg
File image/jpeg, 65k
Caption Figure 14.6 Many households face a stark choice between poaching/land clearing or food insecurity. As an alternative, COMACO helps rural households in Zambia meet their income and nutritional needs through several sustainable income streams. For example, groundnut farmers are provided opportunities to sell their produce at high-value markets, while groundnut shells (a waste product) are pressed into briquettes that can be used as a renewable fuel source.
URL http://books.openedition.org/obp/docannexe/image/10462/img-9.jpg
File image/jpeg, 48k
Caption Figure 14.7 Placing large guard dogs with livestock is a highly effective, non-lethal strategy to reduce predator attacks. The mere presence of the dog is often enough to keep predators away. Several conservation organisations are now providing trained guard dogs to reduce instances of human-wildlife conflicts involving predators.
URL http://books.openedition.org/obp/docannexe/image/10462/img-10.jpg
File image/jpeg, 41k

CC-BY-4.0

The text only may be used under licence CC BY 4.0. All other elements (illustrations, imported files) are “All rights reserved”, unless otherwise stated.

Search OpenEdition Search

You will be redirected to OpenEdition Search