Caatinga Biomes

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Caatinga—South America

José Maria Cardoso da Silva, Department of Geography and Regional Studies, University of Miami, Coral Gables, FL, United States
Thomas E Lacher Jr., Department of Wildlife and Fisheries Sciences, Texas A&M University, College Station, TX, United States
© 2019 Elsevier Inc. All rights reserved.

What is Caatinga? 1
The Landscapes of Caatinga 1
The Biodiversity of Caatinga 4
Threats and Conservation 5
Changing Climatic Conditions 6
The Future of Caatinga 6
References 6

Abstract

The Caatinga is a large semi-arid biome located entirely within the northeastern quadrat of Brazil. It is a region with patterns
of anomalous precipitation, and with occasional multi-year droughts, which has earned it the label the “Polygon of
Drought.” This has affected the adaptations of species within the region as well as the livelihoods of human populations.
The major ecosystem types within the Caatinga biome include open canopy thorn scrub woodlands on shallow soils with an
abundance of cactus, taller semi-arid forests on sedimentary soils, and tall deciduous and semi-deciduous forests on more
nutrient rich substrates. The Caatinga also has a varied topography or small rocky ranges and plateaus. Caatinga biodiversity
was once thought to be species poor with low levels of endemism, but recent research has demonstrated a rich flora and fauna
with high levels of endemism. The Caatinga, however is highly threatened by poor land use practices and expanding large-
scale development, resulting in growing rates of deforestation. This is exacerbated in times of drought, resulting in very low
indicators of human development. Although a number of important protected areas currently exist in the Caatinga, they
represent a small proportion of the biome and are poorly interconnected.

What is Caatinga?

Caatinga is a natural region that lies in the hinterland of northeastern Brazil (Fig. 1). The term caatinga means “white forest” in Tupi,
the language used by the Amerindians that lived in the region (Prado, 2003). Caatinga extends over 900,000 km2 and encompasses
part of the areas of 10 Brazilian states: Piaui, Ceará, Rio Grande do Norte, Paraíba, Pernambuco, Paraíba, Alagoas, Sergipe, Bahia,
and Minas Gerais (Silva et al., 2018a,b). The region is surrounded by the Amazon (on the east), Atlantic forest (on the west), and
Cerrado (on the south). In the north, Caatinga reaches the Atlantic Ocean. As is common among biomes everywhere, Caatinga is
separated from the adjacent terrestrial biomes by transition zones where types of vegetation with distinct ecological requirements
can be found side by side (Eiten, 1972).
Caatinga is one of the most highly distinctive natural regions in South America. Based on the range overlaps of their endemic
species, it has been recognized as a dispersal center (Muller, 1973), an area of endemism (Cracraft, 1985; Haffer, 1985), a
phytogeographic province (Rizzini, 1976), a phytogeographic domain (Queiroz et al., 2018), and a biogeographical province
(Cabrera and Willink, 1980; Udvardy, 1975; Morrone, 2014). Caatinga has also been considered as an ecoregion (Olson et al.,
2011), a biome (IBGE, 2004), or a morphoclimatic domain (Ab’Saber, 1973) because it has a unique combination of physical and
biological features.

The Landscapes of Caatinga

Caatinga is a heterogenous region where seasonally dry tropical forest and woodlands (STDFWs), gallery forests, wetlands, tropical
rain forests, savannas and rupestrian grasslands compose an intricate mosaic (Queiroz et al., 2018). Caatinga’s environmental
heterogeneity is, in turn, a consequence of the interaction of geomorphological, soil, and climate factors (Andrade-Lima, 1981).
Around 71.4% of Caatinga lies on Precambrian crystalline basement, 23.8% on Paleozoic and Mesozoic sedimentary basins, and
4.7% in Quaternary sedimentary basins. Most of Caatinga (42.3%) is formed by peripheral depressions (0–600 m) that are large
flattened or undulating surfaces shaped in both crystalline and sedimentary rocks (Cole, 1986). These depressions are separated by
plateaus (600–2200 m) that are locally known as “serras,” “planaltos,” “tabuleiros,” “chapadas” or “brejos” and that represent residual
formations of old erosion cycles (Cole, 1986). Sometimes, intermediate and mid-altitude surfaces locally known as “patamares” link
the depressions with the plateaus. Regional soils are diverse and several types have been described to the region. The most common
soil classes according to the Brazilian Soil Classification System (Santos et al., 2018) are Neosols (18.7%), Latosols (17.1%),
Argisols (14.1%), and Luvisols (12.8%). Most Caatinga has a semiarid climate, but there are also patches of semi-humid climate due
to orographic effects (Nimer, 1989). The average minimum temperature is 14.8  C (range ¼ 7.6  C–28.1  C), the average mean

Encyclopedia of the World's Biomes https://doi.org/10.1016/B978-0-12-409548-9.11984-0 1


2 Caatinga—South America

Fig. 1 The Caatinga biome is restricted to northeastern Brazil.

temperature is 21.9  C (range ¼ 15.6  C–28.1  C), and the average maximum temperature is 29.7  C (range ¼ 23.0  C–36.3  C).
The average annual rainfall is 1026 mm a year (range ¼ 346 mm/year-1733 mm/year), but 11% of the region receives less than
500 mm/year and 70% receives less than 1000 mm/year. Rainfall is higher from November to April, but it usually is concentrated in
2 or 3 months within this period (Nimer, 1989). The number of dry months varies from six to eleven, with the length of the dry
season declining from the region’s center to its edges (Nimer, 1989). Caatinga has a high interannual variability in rainfall
(sometimes more than 50%), and droughts, defined usually as those years in which the annual rainfall is less than 30% of the
average, can sometimes last for years (Sampaio, 1995). These droughts can have devastating effects on both the ecology of the region
and the human population, and the regions is referred to as the “Polygon of Drought.” Periodic droughts exacerbate the poverty of
the region and are part of the its history, and a frequent theme in the literature and music of northeastern Brazil.
The dominant vegetation in Caatinga is a SDTFW termed locally as “caatingas” (Andrade-Lima, 1981; Prado, 2003) because it
shows a wide variation in structural forms ranging from open cactus scrub (mostly on rock outcrops in the driest areas or areas
deeply transformed by human activities; Fig. 2) to tall, semi-deciduous forests on moister sites with nutrient-rich soils

Fig. 2 General appearance of the Caatinga landscape, dominated by cacti, thorn scrub vegetation.
Caatinga—South America 3

(Andrade-Lima, 1981; Queiroz et al., 2018). From a floristic perspective, there are three subgroups of caatingas: (1) crystalline
caatingas, (2) sedimentary caatingas, and (3) caatinga forests. The crystalline caatingas are the most widespread, as they cover 73%
of the region (Queiroz et al., 2018). They grow on shallow and very stony soils associated with the crystalline basements that do not
retain water for a long time (Fig. 3) and have a flora (Fig. 4) associated with other Neotropical SDTFWs (Queiroz et al., 2018).
In contrast, the sedimentary caatingas grow on deep, nutrient-poor sandy soils of the sedimentary depressions and are found as
islands in three major patches: The Ibiapaba and Arararipe plateaus, the tableland surfaces of the Tucano-Jatobá sedimentary basins,
and the sandy dunes of the São Francisco valley (Queiroz et al., 2018). The flora of sedimentary caatingas seems to be more recent
and associated with other vegetations on sandy soils in eastern South America (Queiroz et al., 2018). These two groups of caatingas
have different phenological patterns, as sedimentary caatingas are not so strongly linked with rainfall patterns, and their floras have
different biogeographic connections with other South American biomes (Queiroz et al., 2018). The caatinga forests is the third and
most distinctive subgroup among caatingas (Prado, 2003). It can be deciduous or semideciduous and is found on nutrient-rich soils
derived from limestone rocks along the valleys of the rivers São Francisco e Gurgéia, on the upland Chapada do Araripe, the Potiguar
basin, and along the slopes of the high plateaus of the Chapada da Diamantina (Prado, 2003). These forests are 20–30 m tall and
dominated by exuberant trees of the genera Tabebuia, Anadenanthera, Myracroduon, Cavanillesia, and Schinopsis (Andrade-Lima, 1981;
Prado, 2003). Some studies suggest that this type of vegetation was once more widespread across the region and that it was replaced
by open physiognomies and woodlands as a consequence of human disturbances (Coimbra-Filho and Câmara, 1996).
In addition to caatingas, patches of other types of vegetation are found in Caatinga and contribute to increase the regional
environmental diversity. For instance, wetlands and seasonally flooded grasslands are found along the main rivers (Siqueira Filho,
2012). Gallery forests with carnauba palm (Copernicia prunifera) are found in the riverbeds near the crystalline caatingas (Queiroz
et al., 2018). Tropical rain forests with strong floristic connections to Atlantic Forest are found along the slopes of Chapada da
Diamantina as well as small and isolated plateaus in the states of Pernambuco, Paraíba, and Ceará (Andrade-Lima, 1982). On the

Fig. 3 Example of the dense thorn scrub understory typical of Caatinga (Photo credit: T. Lacher).

Fig. 4 Cnidoscolus urens, a thorny and spiny member of the family Euphorbiaceae. Many Caatinga plants.
4 Caatinga—South America

top of the Chapada da Diamantina, the Araripe Plateau, and small plateau islands within the region, it is possible to find patches of
cerrados, the tropical xeromorphic savanna that dominate the extensive plateaus of Central Brazil, Bolivia, and Paraguay (Eiten,
1972). Rupestrian grasslands are restricted to the most ancient geomorphological surfaces of the Chapada da Diamantina and are a
vegetation type rich in species and endemism of plant families with unique adaptations that allow them to thrive in very poor soils
(Queiroz et al., 2018).
Caatinga is also a region with unique freshwater ecosystems (Silva et al., 2018a,b). The region has a dense river network
composed of both perennial and intermittent rivers. Perennial rivers are few, and their headwaters are generally outside the Caatinga
(Steffan, 1977). The São Francisco is the largest and most important perennial river in the region, forming a set of unique aquatic
ecosystems along its course. In contrast, intermittent rivers, such as Jaguaribe, are the ones that cease to flow every year or at least
twice every 5 years (Steffan, 1977). They can experience flash floods and prolonged droughts, both of which act as agents of
hydrologic disturbance and have a strong influence on aquatic biological communities (Silva et al., 2018a,b).

The Biodiversity of Caatinga

In the past, the Caatinga biota was thought to be composed mostly of widely distributed species with few being restricted to the
region (Mares et al., 1985). However, recent studies have rejected this hypothesis. In fact, the Caatinga harbors an astounding
diversity of native species with a high proportion of endemic species (Table 1), making it one of the world’s more diverse and
unique tropical drylands (Tabarelli et al., 2018).
In a region with so many endemic species, it is difficult to pinpoint a set of flagship species that can be used as symbols for the
region’s conservation efforts. Nevertheless, Aguiar et al. (2002) proposed two birds and two mammals as flagship species for
Caatinga. The two birds are Lear’s macaw (Anodorhynchus leari) and Spix’s macaw (Cyanopsitta spixii), both of which are endemic to
Caatinga. Lear’s macaw is an impressive metallic blue macaw, superficially similar to the hyacinth macaw (Anodorhynchus
hyacinthinus), but slightly smaller (75 cm long), with a larger head, and lighter in color. This species is known from two colonies
at Toca Velha and Serra Branca, south of Raso da Catarina, Bahia (Collar et al., 2019a). Because of intensive recent conservation
efforts, the population of this species has increased from 140 in 1994 to 1123 in 2010 (Collar et al., 2019a). As a consequence, it has
been downlisted from Critically Endangered to Endangered in the International Union for Nature Conservation (IUCN) Red List.
In contrast with Lear’s macaws, Spix’s macaws seems to be extinct in wild (Collar et al., 2019b). This small blue macaw
(55–57 cm) with pale greenish blue crown and blackish gray bare facial area once lived in the gallery woodlands dominated by
caraíba (Tabebuia caraiba) along the river São Francisco valley (Collar et al., 2019b). However, because the destruction of its habitat
as well as illegal trapping, its population declined fast. When it was rediscovered in nature, its population had only five individuals.
The last known wild individual disappeared in 2000 from near Curacá, Bahia (Sick, 1993). The fate of this species depends on a
successful introduction of a few captive individuals in nature.
The two species of mammals that can be proposed as flagship species are the rock cavy (Kerodon rupestris) and the Brazilian three-
banded armadillo (Tolypeutes tricinctus). The rock cavy is a large (700 g) rodent endemic to Caatinga that is distantly related to
guinea pigs and that occupies rock outcrops where it lives in crevices and forages on foliage (Lacher, 1981; Lacher et al., 1982). With
numerous adaptations to its rocky habitat, the rock cavy has many morphological, ecological, and behavioral similarities to the
hyraxes and gundis of Africa (Mares and Lacher, 1987). It also climbs well and often forages high in trees. When disturbed, it gives a
rapid, trilling alarm call and flees deep into crevices in the boulder piles. The species is heavily hunted but it is considered as least
concern by IUCN Red List (Lacher, 1981).
Different from the rock cavy, the Brazilian three-banded armadillo is considered as vulnerable by the IUCN Red List and its
populations are declining because of heavy hunting. It was once thought to be extinct, but Silva and Oren (1993) rediscovered
populations in tropical deciduous forest remnants in the state of Bahia. Although it has been observed in the Cerrado (Marinho-
Filho et al., 1997; Oliveira 1995), it is largely a Caatinga species. This armadillo has the ability to completely roll itself into a tight
ball as a defensive posture (Silva and Oren, 1993). Because of that behavior, it was selected to be the mascot of the 2014 Soccer
World Cup. In 2015, scientists demanded that the Brazilian government set aside 1000 ha of protected areas in Caatinga for each
goal scored in the tournament (Melo et al., 2014). The Associação Caatinga, an important regional conservation organization,
carries on a program to protect the species.

Table 1 Number of species and proportion of endemics in Caatinga for selected biological groups.

Groups Number of species % Endemic species Source

Flowering Plants 3150 23.0 Queiroz et al. (2018)


Fishes 386 54.1 Lima et al. (2018)
Amphibians 98 20.4 Garda et al. (2018)
Reptiles 79 30.8 Mesquita et al. (2018)
Birds 548 23.0 Araujo and da Silva (2018)
Mammals 183 28.0 Carmignotto and Astúa (2018)
Caatinga—South America 5

The biota of Caatinga has been assembled as a consequence of three biogeographic processes: species production, biotic
exchange, and regional mass extinction (Araujo and da Silva, 2018). Species production is the origin of species within the region,
biotic exchange is the flow of species between regions, and regional mass extinction occurs when many species disappear within a
region as a consequence of regional-scale environmental changes. The relative contribution of each one of these processes seems to
vary across biological groups. Because of the high proportion of endemic species within these groups, species production is more
important in flowering plants (Queiroz et al., 2018), fishes (Lima et al., 2018), amphibians (Garda et al., 2018), and reptiles
(Mesquita et al., 2018). In contrast, because many species living in Caatinga have most of their ranges in adjacent regions, biotic
exchange seems to be more prevalent among birds (Araujo and da Silva, 2018) and mammals (Carmignotto and Astúa, 2018). The
contribution of the regional mass extinction on the present-day composition of the Caatinga’s biota needs still to be investigated,
but recent studies indicate that this process could be important at least among mammals (Castro et al., 2014). In the past, it was
suggested that species endemic to the Caatinga were evolutionarily young and derived from other regions, such as Atlantic Forest or
Chaco (Rizzini, 1976; Andrade-Lima, 1981). However, this hypothesis has been rejected by phylogenetic studies using molecular
markers (Queiroz et al., 2018). The estimated ages of the endemic species of the Caatinga’ range from the early Holocene to Mid-
Miocene, indicating that species production has been continuous within the region (Tabarelli et al., 2018). In addition, Caatinga
endemic species are related to species that live in different South American regions, indicating a complex network of biogeographic
relationships for the region and a very fascinating biogeographic history (Araujo and da Silva, 2018; Queiroz et al., 2018).
As important as an understanding of the biogeographical processes that generated the current Caatinga biota is the identification
of the ecological processes that help to maintain the current biodiversity of the Caatinga. Among the several factors suggested in the
literature, environmental heterogeneity seems to be the most important. A region with high environmental heterogeneity is
expected to maintain more species at local and regional scales than a region with less environmental heterogeneity. The Caatinga
is heterogeneous from an environmental perspective. For instance, Rodrigues e Silva et al. (2000) recognized at least 135 geo-
environmental units and Velloso et al. (2002) recognized nine ecoregions. Although dominated by caatingas, the presence of
enclaves of other vegetation types contributes to sustaining a significantly high number of species across the region. For instance,
among birds, almost 60% of the species require forests (a habitat that covers less than 15% of the region) for their life cycles (Araujo
and da Silva, 2018). The region’s environmental heterogeneity shapes the distribution of the species, forming environmental
gradients. These gradients, in turn, are fundamental in providing mesic refuges and migratory corridors during the long dry season
for several species that are able to move between habitats within landscapes and between landscapes within the region (Silva et al.,
2003). Ecological gradients can also facilitate the speciation among some groups of organisms that have low dispersal, and enable
the evolution and the maintenance of complex and unique types of plant-animal interactions (Leal et al., 2018).

Threats and Conservation

Human activities have been documented within the Caatinga since at least the Late Pleistocene and Early Holocene (Bueno and
Dias, 2015). When the first Europeans disembarked in Brazil in 1500, the indigenous population in the Caatinga was roughly
estimated at 100,000 people (Hemming, 1987). At this time, human impacts on the natural ecosystems of the Caatinga were
possibly restricted to the patches of humid ecosystems found along the few perennial rivers and on the plateau slopes, as crops were
restricted to a few small plots. During the 1600s and 1700s, European colonizers supported by their African slaves slowly colonized
the Caatinga from their coastal settlements. By using the major rivers as routes, they slowly opened up space for settlements that had
cattle ranching and subsistence agriculture as their major economic drivers (Hemming, 1987). This expansion of the colonizers led
to escalating conflicts with the native peoples, who resented their forcible exclusion from ancient hunting and fishing grounds. The
clashes eventually led to the annihilation of the native resistance and the relocation the few survivors to the harshest zones of the
region (Puntoni, 2002). Since then, human settlements have spread out across the whole region and set the basis for the modern
cities. In 2010, the Caatinga was home to 28.6 million people, and the region had some of Brazil’s lowest human development
indicators (Silva et al., 2018a,b). The human footprint can be detected everywhere across the region as traces of pastures, agriculture
fields, roads, fire, reservoirs, cities, and other human infrastructure have accumulated over time as a palimpsest (Silva and
Barbosa, 2018).
Tabarelli et al. (2018) identified three major types of human disturbances in the Caatinga. The first is acute disturbance caused
by the rapid conversion of large areas of native vegetation into human-made ecosystems, dominated by roads, reservoirs, and
commercial agriculture (Silva and Barbosa, 2018). The second type is chronic disturbance caused by the slow but continuous
overexploitation of the native vegetation, such as through establishing slash-and-burn agriculture, collecting firewood, and
browsing by livestock (Albuquerque et al., 2018; Melo, 2018). Finally, the third type consists of the introduction of exotic species
of plants and animals (Almeida et al., 2015). These three types of disturbance together are leading to an irreversible loss of
biodiversity, biomass, and ecological processes that, over time, reduce the provision of critical ecosystem services (e.g., food,
freshwater, and firewood) that local human populations need to survive (Silva et al., 2018a,b). The problem is even larger if one
considers that one-third of the Caatinga has a high potential for desertification and that all attempts to control the expansion of the
existing four desertification nuclei have not been successful (Sá and Angelotti, 2009; Vieira et al., 2015).
Most of the Caatinga currently is covered by anthropogenic rather than natural ecosystems. Silva and Barbosa (2018) combined
the impacts of road expansion, fires, as well as native vegetation loss and estimated that 63.3% of the Caatinga is composed of
anthropogenically altered ecosystems. Human impact is higher in the humid and more productive ecoregions than among those
6 Caatinga—South America

ecoregions where climate and nutrient-poor soils have always constrained human occupation (Velloso et al., 2002). Yet, even the
ecoregions that are less amenable to humans have more than 40% of their areas disturbed. In general, the distribution of the natural
vegetation across the region shows that most of the remnants are becoming small and isolated and that opportunities to protect
large and roadless areas within the Caatinga are quickly becoming scarce (Silva and Barbosa, 2018).
Caatinga is one of the few drylands in the world that has a large-scale science-based conservation plan that was designed together
with stakeholders (Silva et al., 2003) and that has been updated by using new datasets and modern decision-making methods
(Fonseca et al., 2018). In addition, the region has two large and comprehensive conservation initiatives that have been formally
recognized by the Brazilian government: (a) the Caatinga Biosphere Reserve designated in 2001 with 19.9 million hectares and
(b) the Caatinga Ecological Corridor announced in 2006 with 11.7 million hectares. However, despite all this effort, the Caatinga
continues to be poorly protected (Leal et al., 2003). In fact, only 7.4% of the region is within protected areas and most of these
protected areas are not properly funded (Oliveira and Bernard, 2017). If national priorities do not change and more protected areas
are not added to the regional system, the future of the Caatinga would inadequately rely on a few large national parks and ecological
stations that have areas above 50,000 ha. These protected areas are: the National Park Serra das Confusões (823,435 ha), National
Park Boqueirão da Onça (347,557 ha), National Park Chapada Diamantina (152,000 ha), National Park Serra da Capivara
(100,000 ha), Ecological Station Raso da Catarina (99,772 ha), National Park Catimbau (622,300 ha), National Park Cavernas
do Peruaçu (56,800 ha), and State Park Caminho dos Gerais (56,237 ha).

Changing Climatic Conditions

Because of its current interannual rainfall variability, climate is already a major driver of the Caatinga’s socio-ecological systems
(Marengo et al., 2016). The situation is projected to get worse because scenarios of future climate suggest that the near surface air
temperature will increase between 1  C and 4  C and rainfall will reduce by about 0.3 mm day1 by the end of the 21st century
(Torres et al., 2018). In addition, the IPCC (2013) projected with medium to high confidence the following climatic stressors over
the Caatinga: (1) surface soils are projected to dry out; (2) both annual evapotranspiration and runoff are projected to decrease; (3)
days and nights are projected to be warmer; (4) more frequent intense rainfall episodes will be followed by dry and warm periods
without rain; and finally, (5) dry spells are projected to be longer with the possibility of triggering droughts. Climate change
associated with desertification make Caatinga one of the world’s most vulnerable regions to climate change, with potential adverse
impacts on its rich biodiversity, water resources (Andrade et al., 2018), and, consequently, its human population (Torres
et al., 2018).

The Future of Caatinga

What is the future of the Caatinga? It is hard to project, however the current indicators are not promising. Despite some progress in
the last decades, the region still has very low indicators of human development (Silva et al., 2018a,b). Therefore, social inclusion
continues to be a regional issue. From an environmental viewpoint, loss of native ecosystems continues at a fast pace, biodiversity
conservation policies are still poorly elaborated, and all efforts to constrain the desertification frontiers have not been successful.
Projections about the impact of future climate change on the region are grim and the adaptive capacity available at the regional level
to face such changes is limited. A better future for the Caatinga requires the adoption of a development model in which
environmental conservation, social inclusion, economic prosperity and good governance go together (Sachs, 2015). A shift toward
this model demands visionary leaders, persistence, creativity, innovative science and technology, consistent financial and political
support, and a robust and evident connection between the improvement of human livelihoods and the conservation of natural
landscapes (Leal et al., 2005). These basic enabling conditions for change have never been in place during the last several decades
and, unfortunately, are unlikely to be put in place in the near future.

References

Ab’Saber AN (1973) O domí nio morfoclimático semi-árido das caatingas brasileiras. Geomorfologia 43: 1–39.
Aguiar J, Lacher T, and Silva JMC (2002) The Caatinga. In: Mittermeier RA, Mittermeier CG, and Robles Gil P, et al. (eds.) Wilderness—Earth’s Last Wild Places, pp. 175–180.
Mexico: Cemex.
Albuquerque UP, Araújo EL, de Castro CC, et al. (2018) People and natural resources in the Caatinga. In: Caatinga: The largest tropical dry forest region in South America,
pp. 303–333. Springer International Publishing.
Almeida WR, Lopes AV, Tabarelli M, and Leal IT (2015) The alien flora of Brazilian Caatinga: Deliberate introductions expand the contingent of potential invaders. Biological Invasions
17: 51–56.
Andrade EM, Aquino DN, Chaves LCG, et al. (2018) Water as capital and its uses in the Caatinga. In: Caatinga: The largest tropical dry forest region in South America, pp. 281–302.
Springer International Publishing.
Andrade-Lima D (1981) The Caatinga dominium. Revista Brasileira de Botânica 4: 149–153.
Andrade-Lima D (1982) Present-day forest refuges in northeastern Brazil. In: Prance GT (ed.) Biological diversification in the tropics, pp. 245–251. Nova York: Columbia University
Press.
Caatinga—South America 7

Araujo HFP and da Silva JMC (2018) The avifauna of the Caatinga: Biogeography, ecology, and conservation. In: Caatinga: The largest tropical dry forest region in South America,
pp. 181–210. Springer International Publishing.
Bueno L and Dias A (2015) Povoamento inicial da América do Sul: Contribuições do contexto brasileiro. Estudos Avançados 29(83): 119–147.
Cabrera ÁL and Willink A (1980) Biogeografí a de América Latina. Washington, D.C.: OEA: Secretaría General de la Organización de los Estados Americanos, Programa Regional de
Desarrollo Científico y Tecnológico.
Carmignotto AP and Astúa D (2018) Mammals of the Caatinga: Diversity, ecology, biogeography, and conservation. In: Caatinga: The largest tropical dry forest region in South America,
pp. 211–254. Springer International Publishing.
Castro MC, Montefeltro FC, and Langer MC (2014) The quaternary vertebrate fauna of the limestone cave Gruta do Ioiô, northeastern Brazil. Quaternary International 352: 164–175.
Coimbra-Filho AF and Câmara IG (1996) Os Limites Originais Do Bioma Mata Atlantica Na Regiao Nordeste Do Brasil. Fundação Brasileira para a Conservação da Natureza.
Cole MM (1986) The savannas: Biogeography and Geobotany. London: Academic Press.
Collar N, Boesman P, Sharpe CJ, Kirwan GM, and Garcia EFJ (2019a) Lear’s Macaw (Anodorhynchus leari). In: del Hoyo J, Elliott A, Sargatal J, Christie DA, and de Juana E (eds.)
Handbook of the Birds of the World Alive. Barcelona: Lynx Edicions retrieved from https://www.hbw.com/node/54614 on 12 June 2019.
Collar N, Boesman P, Sharpe CJ, and Kirwan GM (2019b) Spix’s Macaw (Cyanopsitta spixii). In: del Hoyo J, Elliott A, Sargatal J, Christie DA, and de Juana E (eds.) Handbook of the
Birds of the World Alive. Barcelona: Lynx Edicions retrieved from https://www.hbw.com/node/54615 on 12 June 2019.
Cracraft J (1985) Historical biogeography and patterns of differentiation within the South American avifauna: Areas of endemism. Ornithological Monographs (36): 49–84.
Eiten G (1972) The cerrado vegetation of Brazil. The Botanical Review 38(2): 201–341.
Fonseca CR, Antongiovanni M, Matsumoto M, et al. (2018) Conservation opportunities in the Caatinga. In: Caatinga: The largest tropical dry forest region in South America,
pp. 429–443. Springer International Publishing.
Garda AA, Stein MG, Machado RB, et al. (2018) Ecology, biogeography, and conservation of amphibians of the Caatinga. In: Caatinga: The largest tropical dry forest region in South
America, pp. 133–149. Springer International Publishing.
Haffer J (1985) Avian zoogeography of the Neotropical lowlands. Ornithological Monographs (36): 113–146.
Hemming J (1987) Indians and the frontier. In: Bethell L (ed.) Colonial Brazil, pp. 145–189. Cambridge: Cambridge University Press.
IBGE (2004) Mapa de biomas do Brasil. Rio de Janeiro: Instituto Brasileiro de Geografia e Estatística. Available at: ftp://ftp.ibge.gov.br/cartas_e_mapas/Mapas_e_Murais/.
IPCC (2013) Climate change 2013: The physical science basis. In: Stocker TF, Qin D, Plattner G-K, Tignor M, Allen SK, Boschung J, Nauels A, Xia Y, Bex V, and Midgley PM (eds.)
Contribution of working group I to the fifth assessment report of the intergovernmental panel on climate change. Cambridge, United Kingdom and New York, NY, USA: Cambridge
University Press.
Lacher TE (1981) The comparative social behavior of Kerodon rupestris and Galea spixii and the evolution of behavior in the Caviidae. Bulletin of Carnegie Museum of National History
17: 1–71.
Lacher TE Jr., Willig MR, and Mares MA (1982) Food preference as a function of resource abundance with multiple prey types: An experimental analysis of optimal foraging theory.
American Naturalist 120: 297–316.
Leal IR, Tabarelli M, and Silva JMC (2003) Ecologia e conservação da Caatinga. Recife: Editora Universitária da UFPE.
Leal IR, Silva JMC, Tabarelli M, and Lacher TE Jr. (2005) Changing the course of biodiversity conservation in the caatinga of northeastern Brazil. Conservation Biology 19: 701–706.
Leal IR, Lopes AV, Machado IC, et al. (2018) Plant-animal interactions in the Caatinga: Overview and perspectives. In: Caatinga: The largest tropical dry forest region in South America,
pp. 255–278. Springer International Publishing.
Lima SMQ, Ramos TPA, da Silva MJ, et al. (2018) Diversity, distribution, and conservation of the Caatinga fishes: Advances and challenges. In: Caatinga: The largest tropical dry forest
region in South America, pp. 97–131. Springer International Publishing.
Marengo JA, Torres RR, and Alves LM (2016) Drought in Northeast Brazil—Past, present, and future. Theoretical and Applied Climatology 129: 1189–1200.
Mares MA and Lacher TE (1987) Ecological, morphological, and behavioral convergence in rock-dwelling mammals. In: Genoways HH (ed.) Current Mammalogy, pp. 307–348. Boston,
MA: Springer.
Mares MA, Willig MR, and Lacher TE (1985) The Brazilian Caatinga in south American zoogeography: Tropical mammals in a dry region. Journal of Biogeography 12(1): 57–69. Wiley.
Marinho-Filho J, Guimarães M, Reis ML, Rodrigues F, Torres O, and Almeida G (1997) The discovery of the Brazilian three-banded armadillo in the Cerrado of Central Brazil. Edentata
3: 11–13.
Melo FPL (2018) The socio-ecology of the Caatinga: Understanding how natural resource use shapes an ecosystem. In: Caatinga: The largest tropical dry Forest region in South
America, pp. 369–382. Springer International Publishing.
Melo FP, Siqueira JA, Santos BA, et al. (2014) Football and biodiversity conservation: FIFA and Brazil can still hit a green goal. Biotropica 46(3): 257–259.
Mesquita DO, Costa GC, Garda AA, et al. (2018) Species composition, biogeography, and conservation of the Caatinga lizards. In: Caatinga: The largest tropical dry forest region in
South America, pp. 151–180. Springer International Publishing.
Morrone JJ (2014) Biogeographical regionalisation of the Neotropical region. Zootaxa 3782(1): 1–110.
Muller P (1973) Dispersal centers of terrestrial vertebrates in the Neotropical. Biogeographica 2: 1–244.
Nimer E (1989) Climatologia do Brasil. Rio de Janeiro: IBGE, Departamento de Recursos Naturais e Estudos Ambientais.
Oliveira TG (1995) The Brazilian three-banded armadillo Tolypeutes tricinctus in Maranhão. Edentata 2: 18–19.
Oliveira APC and Bernard E (2017) The financial needs vs. the realities of in situ conservation: An analysis of federal funding for protected areas in Brazil’s Caatinga. Biotropica
49: 745–752.
Olson DM, Dinerstein E, Wikramanayake ED, et al. (2011) Terrestrial ecoregions of the world: A new map of life on earth. Bioscience 51(11): 933–938.
Prado D (2003) As caatingas da América do Sul. In: Leal IR, Tabarelli M, and Silva JMC (eds.) Ecologia e conservação da Caatinga, pp. 3–73. Recife: Editora Universitária da UFPE.
Puntoni P (2002) A Guerra dos Bárbaros: povos indí genas e a colonização do sertão nordeste do Brasil 1650–1720. São Paulo: Hucitec, Universidade de São Paulo, and FAPESP.
Queiroz LP, Cardoso D, Fernandes MF, et al. (2018) Diversity and evolution of flowering plants of the Caatinga domain. In: Caatinga: The largest tropical dry forest region in South
America, pp. 23–63. Springer International Publishing.
Rizzini CT (1976) Tratado de fitogeografia do Brasil—Aspectos sociológicos e florí sticos. São Paulo: Editora de Humanismo, Ciência e Tecnologia.
Rodrigues e Silva FB, Santos JCP, Souza Neto NC, Silva AB, Riché GR, Tonneau JP, Brito LTL, Correia RC, Silva FHBB, Silva CP, Leite AP, Oliveira Neto MB, Parahyba RBV, Araújo
Filho JC, Cavalcanti AC, Burgos N, and Reis RMG (2000) Zoneamento Agroecológico do Nordeste: Diagnóstico e Prognóstico. Recife and Petrolina: CD-ROM. Embrapa Solos
E Embrapa Semi-Árido.
Sá IB and Angelotti F (2009) Degradação ambiental e desertificação no semi-árido brasileiro. In: Angelotti F, Sá IB, Menezes EA, and Pellegrino GQ (eds.) Mudanças climáticas e
desertificação no semi-árido brasileiro, pp. 59–88. Petrolina: Embrapa Semiárdo.
Sachs JD (2015) The age of sustainable development. New York: Columbia University Press.
Sampaio EVSB (1995) Overview of the Brazilian Caatinga. In: Bullock SH, Mooney HA, and Medina E (eds.) Seasonally dry forests, pp. 35–58. Cambridge: Cambridge University Press.
Santos HG, Jacomine PKT, Anjos LH, et al. (2018) Sistema brasileiro de classificação de solos. Brasília: ENBRAPA.
Sick H (1993) Birds in Brazil. A natural history. Princeton, New Jersey: Princeton University Press.
Silva JMC and Barbosa LCF (2018) Impact of human activities on the Caatinga. In: Caatinga: The largest tropical dry forest region in South America, pp. 359–368. Springer
International Publishing.
Silva JMC and Oren DC (1993) Observations on the habitat and distribution of the Brazilian three-banded armadillo Tolypeutes tricinctus, a threatened Caatinga endemic. Mammalia
57(1): 149–152.
8 Caatinga—South America

Silva JMC, Souza MA, Bieber AGD, and Carlos CJ (2003) Aves da Caatinga: Status, uso do habitat e sensitividade. In: Leal IR, Tabarelli M, and Silva JMC (eds.) Ecologia e conservação
da caatinga, pp. 237–273. Recife: Editora Universitária.
Silva JMC, Barbosa LCF, Leal IR, et al. (2018a) The Caatinga: Understanding the challenges. In: Caatinga: The largest tropical dry forest region in South America, pp. 3–19. Springer
International Publishing.
Silva JMC, Barbosa LCF, Pinto LPS, et al. (2018b) Sustainable development in the Caatinga. In: Caatinga: The largest tropical dry forest region in South America, pp. 445–458.
Springer International Publishing.
Siqueira Filho JA (2012) Flora das Caatingas do Rio São Francisco: História natural e conservação. Rio de Janeiro: Andrea Jakobson Estúdio.
Steffan ER (1977) Hidrografia. In: IBGE (ed) Geografia do Brasil: Região Nordeste. IBGE, Rio de Janeiro, p. 111–133.
Tabarelli M, Leal IR, Scarano FR, et al. (2018) The future of the Caatinga. In: Silva JMC, Leal IR, and Tabarelli M (eds.) Caatinga: The largest tropical dry forest region in South America,
pp. 461–474. Springer.
Torres RR, Lapola DM, and Gamarra NLR (2018) Future climate change in the Caatinga. In: Caatinga: The largest tropical dry Forest region in South America, pp. 383–410. Springer
International Publishing.
Udvardy MDF (1975) A classification of the biogeographical provinces of the world. In: IUCN Occasional Paper 18, pp. 1–48. Morges, Switzerland: IUCN.
Velloso AL, Sampaio EVSB, and Pareyn FGC (2002) Ecorregiões propostas para o Bioma Caatinga. Recife: Associação Plantas do Nordeste—Instituto de Conservação Ambiental—The
Nature Conservancy do Brasil.
Vieira RMSP, Tomasella J, Alvalá RCS, Sestini MF, Affonso AG, Rodriguez DA, Barbosa AA, Cunha APMA, Valles GF, Crepani E, De Oliveira SBP, De Souza MSB, Calil PM, De
Carvalho MA, Valeriano DM, Campello FCB, and Santana MO (2015) Identifying areas susceptible to desertification in the Brazilian northeast. Solid Earth 6: 347–360.

You might also like