You are on page 1of 50

Conservation in the Amazon: Evolution and Situation

Conservation in the Amazon: Evolution and Situation  


Marc Dourojeanni
Subject: Environmental Processes and Systems, Environmental History
Online Publication Date: Apr 2019 DOI: 10.1093/acrefore/9780199389414.013.41

Summary and Keywords

In 1945 the Amazon biome was almost intact. Marks of ancient cultural developments in
Andean and lowland Amazon had cicatrized and the impacts of rubber and more recent
resources exploitation were reversible. Very few roads existed, and only on the Amazon’s
periphery. However, from the 1950s, but especially in the 1960s, Brazil and some Andean
countries launched ambitious road-building and colonization processes. Amazon occupa­
tion heavily intensified in the 1970s when forest losses began to raise worldwide concern.
More roads continued to be built at a geometrically growing pace in every following
decade, multiplying correlated deforestation and forest degradation. A no-return point
was reached when interoceanic roads crossed the Brazilian-Andean border in the 2000s,
exposing remaining safe havens for indigenous people and nature. It is commonly esti­
mated that today no less than 18% of the forest has been substituted by agriculture and
that over 60% of that remaining has been significantly degraded.

Theories regarding the importance of biogeochemical cycles have been developed since
the 1970s. The confirmation of the role of the Amazon as a carbon sink added some inter­
national pressure for its protection. But, in general, the many scientific discoveries re­
garding the Amazon have not helped to improve its conservation. Instead, a combination
of new agricultural technologies, anthropocentric philosophies, and economic changes
strongly promoted forest clearing.

Since the 1980s and as of today Amazon conservation efforts have been increasingly di­
versified, covering five theoretically complementary strategies: (a) more, larger, and bet­
ter-managed protected areas; (b) more and larger indigenous territories; (c) a series of
“sustainable-use” options such as “community-based conservation,” sustainable forestry,
and agroforestry; (d) financing of conservation through debt swaps and climate change’s
related financial mechanisms; and (e) better legislation and monitoring. Only five small
protected areas have existed in the Amazon since the early 1960s but, responding to the
road-building boom of the 1970s, several larger patches aiming at conserving viable sam­
ples of biological diversity were set aside, principally in Brazil and Peru. Today around
22% of the Amazon is protected but almost half of such areas correspond to categories
that allow human presence and resources exploitation, and there is no effective manage­
ment. Another 28% or more pertains to indigenous people who may or may not conserve
the forest. Both types of areas together cover over 45% of the Amazon. None of the
Page 1 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

strategies, either alone or in conjunction, have fully achieved their objectives, while de­
velopment pressures and threats multiply as roads and deforestation continue relentless­
ly, with increasing funding by multilateral and national banks and due to the influence of
transnational enterprises.

The future is likely to see unprecedented agriculture expansion and corresponding inten­
sification of deforestation and forest degradation even in protected areas and indigenous
land. Additionally, the upper portion of the Amazon basin will be impacted by new, larger
hydraulic works. Mining, formal as well as illegal, will increase and spread. Policymakers
of Amazon countries still view the region as an area in which to expand conventional de­
velopment while the South American population continues to be mostly indifferent to
Amazon conservation.

Keywords: Amazon, pre-Columbian occupation, deforestation, forest degradation, scientific support, protected ar­
eas, indigenous land, forest management, community development, policy and legislation for conservation

Introduction
This is an overview of the evolution of the environmental situation in the Amazon since
the end of World War II (1945) when development in the region was incipient. Three as­
pects are discussed: (a) the historical and current impact of development on forests in
terms of deforestation and forest degradation, (b) the influence of the progress of science
and technology on conservation measures, and (c) the diversity, scope, and effectiveness
of conservation measures.

Many criteria are available to describe the Amazon. For this article the concept of biome
is applied (Whittaker, 1962). The Amazon biome is grossly equivalent to the Amazon for­
est, originally demarcated by the tree line of the cloud forests mostly at 3,600–3,800 me­
ters in the eastern Andean flanks of Bolivia and Peru and lower in Ecuador and Colombia
and the Atlantic Ocean, covering some 640 million hectares of Brazil (67.8% of Amazon),
Peru (13%), Bolivia (11.2%), Colombia (5.5%), Ecuador, Guyana, Venezuela, Surinam, and
French Guyana (Commission on Development and Environment for Amazonia [CDEA],
1992). It includes two main basins, Amazon and Orinoco, which cover some 7 million
square kilometers and a great diversity of ecosystems, determined by altitude in the An­
dean slopes and by orography, soil, and climatic conditions everywhere, imposing a great
biological diversity (Holdridge, 1947; Prance, 1982; Udvardy, 1975).

The presence or absence of forest is the most obvious and useful indicator of the environ­
mental condition of Amazon ecosystems. Deforestation has been measured for decades in
the Amazon and, despite information that is being politicized, it offers a good picture of
the status of conservation. Forest degradation is also important, but its measurement and
interpretation are much more difficult to achieve (Lund, 2009).

Page 2 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

The term “Amazon conservation” used in this article is equivalent to the concept of sus­
tainable development, meaning efforts made in the region to harmonize economic devel­
opment, social progress, and environmental protection (World Commission on Environ­
ment and Development [WCED], 1987). It implies limited well-justified deforestation, no
irreversible degradation nor waste of natural resources, and the proficiency of conserva­
tion measures such as protected areas. This article is intended to offer a measure of the
gap between what could be considered sustainable development and the changing reality
over the decades.

Historic Occupation and Its Impacts


Pre-Columbian Human Presence

The Amazon has had a human population possibly as long as any other biome in South
America (Mann, 2011; Roosevelt, 2013). Early Andean cultural developments such as
Chavin, located out of the Amazon biome but within the Amazon basin, gave ground to
the theory of an Amazon origin for all ancient Peruvian cultures (Tello, 1960). Caral dis­
coveries (Shady, 2006) among others in the Peruvian Pacific coast challenged this theory
as they are older than Chavin. However, recent findings demonstrated a civilization prob­
ably even older than Caral was situated in Jaen (Peru), the easiest contact location be­
tween the Amazon and the Pacific Coast (Atwood, 2011, 2017; Clasby & Meneses, 2012).
Thus, the discussion is still open. Nonetheless, humans have existed in the Amazon for
millennia, as long ago as 12,000 years BCE in Peru (Church, 1994). While previous views
considered the Amazon’s human carrying capacity limited (Meggers, 1971) it is now as­
sumed that before arrival of Europeans, in one or more moments of history, the total pop­
ulation of the Amazon could have reached several million inhabitants (Denevan, 1992,
2003; Newson, 1996).

Overall, four types of societies probably occurred: (a) civilizations that left large engi­
neering works, mostly located in the western and southern Amazon peripheries; (b) civi­
lizations with important human concentrations located along the great rivers and in the
Amazon delta; (c) forest villagers who practiced agriculture, including those associated
with the “black earth” sites; and (d) forest villagers who did not practice agriculture, in­
cluding hunters, fishermen, and itinerant gatherers. Obviously, there is a complex gradi­
ent in time and space between these peoples and it is not always possible to differentiate
them completely.

Part of the first group settled in the high humid jungles of Peru, Ecuador, and Colombia,
built stone cities, such as Pajatén in Peru or Pastaza and Sangay in Ecuador (Lumbreras,
1981, 2013), and developed agricultural practices similar to those used in the Andes.
They deforested important areas and cultivated corn, among other adapted Andean
plants, as well as many useful plants known to locally preexisting peoples (Brack, 2003;
León, 2013). In the Amazonian southwest, in the Pampas de Mojos in the Bolivian Beni,
another civilization built impressive hydraulic infrastructures. They also grew corn and

Page 3 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

squash, among other local plants (Denevan, 1970, 1976). And in Acre (Brazil), a consider­
able number of earth structures (geoglyphs) have been discovered that reveal a civiliza­
tion that may or may not be related to the previous one but that also developed signifi­
cant agriculture (Watling et al., 2017). Hydraulic works like those mentioned, although of
lesser proportion, existed in several other places in the Amazon (Pärssinen, Schaan, &
Ranzi, 2009; Ranzi & Aguiar, 2004). But there is much more evidence of widespread pre-
Columbian cultural presence in the Amazon as well as of exchanges with other cultures
far beyond the Amazon biome (Lathrap, 1973). Most of these civilizations had already dis­
appeared or were decadent when the Europeans arrived. Some of the Amazonian west
was, by then, occupied by the Incas, who had also built new citadels such as Machu Pic­
chu and Choquequirao farther south (Burga, 2008).

Amazonian civilizations obviously used fire to supplement their stone and in some cases
bronze tools when clearing trees and expanding agriculture. They may have built terraces
to limit erosion. Fire was also used to open up the forests that covered the land where the
great hydraulic works of the Beni were built in Bolivia and elsewhere, and its use has
been demonstrated in the case of the civilization that built the Acre geoglyphs in Brazil.
These works, like any infrastructure that modifies the natural drainage, caused major en­
vironmental alterations whose impacts are visible even today, changing the floristic com­
position and therefore the entire trophic chain (Watling et al., 2017). It is not known how
many hectares these diverse and large cultural developments in different parts of the
Amazon may have deforested over time but, considering the extent occupied by them, the
area may have been very significant.

The second group corresponds to the civilizations that occupied the banks of the Amazon­
ian rivers and took advantage of the várzeas or floodplains. These were reported by Orel­
lana during his expedition of the Amazon and by the Portuguese who sailed from the delta
and encountered the Marajo cultures (Roosevelt, 2000). Everything indicates that they
had relatively large inhabited centers with complex societies and a significant population,
with agriculture fueled by the annual renewal of soil fertility, which did not require signif­
icant deforestation. They also took advantage of the alluvial forests and nearby highland
forests for hunting, in addition to fishing, which was obviously the main source of animal
protein (Blatrix et al., 2018; Mashuta, Ovando, & McKey, 2018). Possibly these riparian
cultures were the bulk of the pre-Columbian Amazonian population. They were almost
completely decimated by diseases brought by Spaniards and Portuguese at initial contact
(Hamilton, Walker, & Kesler, 2014). Those civilizations left less traces because of their ex­
posure to exceptional floods that literally washed away many archaeological sites.

The other two groups were unlikely to have been very different from today’s indigenous
people. They developed small settlements, relatively isolated, that formed nations with
differentiated territories. They developed a diversified agriculture and may have altered
the nearby forests by propagating useful tree species (Crystal, McMichael, Matthews-
Bird, Farfan-Rios, & Feeley, 2017; Levis et al., 2017; McMichael, Matthews-Bird, Farfan-
Rios, & Feeley, 2017). Some of them seem to have created the many sites known as terra
preta or black earth to ensure lasting soil fertility (Basso & Kimura, 2006; Costa, Kern,

Page 4 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Pinto, Eleotério, & Souza, 2004; Erickson, 2003). Considering the extent of the occupa­
tion of these indigenous people in the Amazon, confirmed by satellite imagery and
drones, it is evident that they caused some deforestation to establish agriculture. A 3.2%
figure has been mentioned for terra preta lowlands (Wade, 2014). Others practiced typical
migratory agriculture, or slash and burn, rotating the cultivation areas. And, finally, there
were also indigenous groups that depended principally on hunting, fishing, and gather­
ing. A few small groups of uncontacted indigenous groups still live in remote parts of the
region (Brackelaire, 2006).

Access to the Amazon was a lot easier by navigation up the Amazon River than by land
down the Andes. Thus, despite precious woods such as mahogany being extracted, some
plantations, such as coca, were established in the upper Amazon near Spanish cities;
some other forest products were extracted by the Portuguese using the city of Belem do
Para (founded in 1616) as a base. The search for goods, mostly gold and precious stones,
allowed the establishment of a network of new small villages along the Amazon River and
its main tributaries that developed subsistence agriculture. It was only by the mid-19th
century, when the rubber boom began, that some newcomers considered the establish­
ment of land properties that later took the form of farms, when the natural rubber lost its
importance. Despite all the evils related to the two phases of the rubber boom, it has not
caused obvious permanent damage to forests (Bunker, 1984). Simultaneously, and later,
species such as river turtles, rosewood, Brazil nuts, as well as giant otter, ocelot and
jaguar skins, and peccaries and caiman hides, among others, were also subject to heavy
exploitation, but none has been a proven cause of biological extinction nor caused any
significant deforestation. Between the two World Wars a small number of short non-paved
roads were built, linking some Andean cities to the upper Amazon, while some farms for
tea, sugar cane, cacao, and coffee began to be developed. A similar situation was develop­
ing in the southern portion of the Brazilian Pará state.

In summary, despite evidence of a very old and quite significant agricultural occupation of
the Amazon that unquestionably has made perdurable ecological changes in large tracks
of the region, most of these impacts were mitigated by a new relatively natural climax
caused by a population gap most probably induced by diseases and disarray following Eu­
ropean arrival. By the end of World War II, despite heavy and irrational exploitation of
natural resources in the previous two centuries, deforestation of the Amazon was not sig­
nificant. While not measured, forest degradation caused by unrestricted hunting and ex­
traction of timber and non-timber products was at that time probably more significant
than deforestation.

Postwar Development in the Amazon

The last years of the 1940s and especially the 1950s witnessed the beginning of the boom
of road-building in most Amazon countries. The biggest push in that period was in Peru
and Ecuador, and also in Bolivia. In these countries, so-called “penetration” roads were
intended to reach upper Amazon valleys and, especially, eastern navigable rivers. Brazil
was still relying on river navigation, but in the late 1950s everything changed as its long-

Page 5 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

standing policy of occupation of the west materialized with the move of the capital city
from Rio de Janeiro to Brasilia (inaugurated in 1960) in the center of the country (Martins
de Souza, 1996). A first step has been the Brasilia–Belem road (BR-010). The big push
came in the early 1970s with the construction of the 4,000-kilometer-long Trans-Amazo­
nia Highway (BR-230) linking the Brazilian Amazon states from the Atlantic to the bor­
ders of Colombia, Ecuador, and Peru (Fearnside, 2005).

During the mid-1970s another road (BR-364) was built to link Cuiabá, Mato Grosso’s capi­
tal, to Porto Velho in Rondônia and Rio Branco in Acre. In the 1980s this road was paved
with financing from the World Bank (in Mato Grosso and Rondonia) and the Inter-Ameri­
can Development Bank (in Acre). This road attracted more people more quickly than the
Trans-Amazonia road, inducing the government to launch the Northwest Development
Program (POLONOROESTE), with additional investments from these banks aiming at ra­
tionalizing land occupation and establishing sustainable settlements, protected areas, and
indigenous reserves. This project, as others later, had only limited success as uncon­
trolled deforestation and mining proliferated (Dourojeanni, 1985A; Fearnside, 1989; Mil­
likan, 1984).

Several other important roads were also initiated in the 1970s linking the states of Mato
Grosso and Para, Tocantins and Para (BR-320), and the states of Amazonas and Roraima.
The roads between Cuiaba and Santarem (BR-163) and between Manaus and Boa Vista
(BR-174) to the Venezuelan frontier (Barni, Fearnside, & Graça, 2015; Fearnside, 2007;
Rodrigues & da Silva Pinheiro, 2011) were inaugurated in 1990. Several more have been
built or are under construction, including the opening of a link between the state of Ama­
zonas and the network of roads in Rondônia and Acre (BR-319) and another road opening
the Xingu Valley. Many thousands of kilometers of regional roads complement and amplify
this basic network.

Concerned about Brazilian geopolitics, neighboring countries adopted equivalent mea­


sures regarding the Amazon (Medina, 1980). They also built ambitious roads aiming to
link urban centers with lowland navigable rivers, such as the Lima–Pucallpa (Ucayali Riv­
er) road. Another important Peruvian effort has been the Marginal de la Selva (PE 5N), a
road that opened to colonization huge areas of the Huallaga Valley in the 1960s, and that
was expanded in the 1980s and later. Peru also adopted the policy of “living frontiers,” es­
tablishing precarious jungle settlements near the limits of Brazilian towns (Belaunde,
1969; Dourojeanni, 1990). In Bolivia and Ecuador, roads in the Amazon were and still are
often more related to oil exploitation than to colonization, but the end result has been the
same, as they have attracted peasants, loggers, and gold diggers (Rudel, 1983). However,
these countries also adopted the concept of the Marginal (known as Troncal de la Selva
or E 45 in Ecuador), intending to replicate in the upper Amazon the concept of the mainly
coastal Pan-American Highway.

The 1990s saw a constant amplification of the Amazon road network and improvement of
existing routes, including bridge-building and paving. Main roads as well as rivers were
interlinked with new roads. While the main roads or axes were federally or nationally

Page 6 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

planned decisions, most secondary and all tertiary roads were the result of provincial, lo­
cal, or even individual decisions, frequently built only to facilitate logging, ranching, or
mining (Dourojeanni, 2011A; Perz, Overdevest, Caldas, Walker, & Arima, 2007). These un­
planned and often illegal roads were usually adopted and improved by the ministries of
public transportation.

The 21st century saw the first road connecting the Brazilian road network with those of
the Andean countries. The interoceanic roads, sponsored by Brazil through the Integra­
tion of Regional Infrastructure in Latin America Initiative (IIRSA), opened to resources
exploitation a vast region of the Amazon frontier between Brazil and Peru that had previ­
ously remained a haven for indigenous people and wildlife (van Dijc, 2013). The region of
Madre de Dios, in Peru, has become easily accessible for agriculture, logging, and espe­
cially gold digging since 2011, when the interoceanic road (Interoceánica Sur) between
Rio Branco in Brazil and Cuzco in Peru (PE 26B) became fully paved and operative (Alber­
ti & Pereyra, 2018). Its enormous negative impact on forests and local society has been
predicted (Dourojeanni, 2006) and confirmed (Fernandez, 2010; Webster & Haviv, 2012).
Another interoceanic road (Interoceánica Central) is currently advancing between Pucall­
pa in Peru and the Brazilian road system at Cruzeiro do Sul (Glave et al., 2012), while a
bimodal (road and waterway) transport system (Interoceánica Norte) is already in place
in northern Peru ensuring a connection with Manaus. A branch of this main road will also
connect the Amazon regions of Ecuador and Peru (Dourojeanni, 2011A, 2013B). There is
still a large forest patch unimpacted by roads in the Brazilian state of Amazonas and its
nearby Peruvian department of Loreto.

The Ecuadorian Amazon region is currently well served by roads that connect with the
Peruvian system and more are being planned, including with Brazil through the bimodal
Manta–Manaus project that includes roads and waterways (Izko, 2012). While still rela­
tively free of roads, Colombia has plans to build more roads in the Amazon and to trans­
form rivers into waterways (Arenas, Zúñiga, & Mayordomo, 2011). Venezuela and French
Guyana are today connected with Brazil. Guyana and Surinam roads are mostly associat­
ed with mining.

Road expansion plans are aggressively being developed everywhere in the Amazon. Be­
tween 2004 and 2007 around 17,000 km of roads per year were built in the Brazilian
Amazon region alone (Ahmed, Souza, Ribeiro, & Ewers, 2013). Around 100,000 km of
roads, of which 64.5% are not paved, were identified in the Amazon in 2012 (Red
Amazónica de Información Socioambiental Georreferenciada [RAISG], 2012). In 2013 the
Brazilian government registered 178,000 km of roads in its Amazon region, including
those under imminent construction. In Peru a new road from Yurimaguas is advancing to
the Marañon River and several others are making progress from the city of Pucallpa to
the north and from the Ucayali and Amazon Rivers to the Brazilian frontier (Finer & Villa,
2018). The still isolated city of Iquitos will soon be connected to the Interoceanica Norte.
Bolivia is also linking its Pando region with the Interoceanica Sur in Peru. The economic

Page 7 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

justification of these roads is not always evident, especially when considering environ­
mental costs (Malky, Reid, Ledezma, & Fleck, 2011).

Roads are supplemented by the construction of waterways and railways. These are less
environmentally damaging and may be better options if they replace roads. However, in
the Amazon these public works are scarce and poorly maintained. In addition, especially
in Peru, they are often parallel routes to existing or planned roads (Dourojeanni, 2013B).
The main Amazon Peruvian rivers are currently the subject of a Chinese investment
project to improve their navigability in order to better connect with the Brazilian Amazon
waterways (ProInversión, 2016). Several international railways are being proposed in
Brazil to the Pacific coast in addition to interoceanic roads. The two most advanced pro­
posals would link the Brazilian railways system with ports in the Peruvian northern coast
(Caillaux, Novak, & Ruiz, 2016) while another, also serving Bolivia, is planned to go to
ports on the southern Peruvian coast (Small, 2017). Both are said to be financed by China.

Brazil built the first large dams (Tucurui and Balbina) in the Amazon lowlands in the
1980s. By then, some smaller or mid-sized dams already existed in the higher watershed,
in the Andean countries, and in the Orinoco basin, but these, with few exceptions, had
limited negative impact. Unfortunately, the lowlands’ huge reservoirs proved to be very
harmful in terms of loss of natural ecosystems, increase in greenhouse gas emissions,
blockage of fish migration, and creation of anoxic water, as well as producing methane
and providing conditions for methylation of mercury (Fearnside, 2001, 2006). The grow­
ing energy requirements of Brazil in the 1990s and 2000s justified the building of many
more dams covering almost every river of the Amazon. Several other very large dams are
being built in Brazil, such as Belo Monte in the Xingu River and Jirau and Santo Antonio
in the Madeira River (Fearnside, 2012). Brazil is currently building or planning to reach a
production of 42,529.5 MW in its Amazon region (Little, 2013), and it is also putting polit­
ical pressure on its Andean neighbors, especially Peru and Bolivia, to build dams in their
physiographically more advantageous Amazon territory. Fifteen large dams are consid­
ered a priority inside the Peruvian territory (Serra, 2010). The impacts of dams that are
located out of the Amazon biome but in its basin are also very serious in terms of disrup­
tion of the ecological connectivity of the Amazon River to the Andes, causing substantial
impacts on nutrient cycling and fish populations. Of the 413 dams already in operation,
under construction, or proposed, 256 are in Brazil, 77 in Peru, 54 in Ecuador, 14 in Bo­
livia, six in Venezuela, two in Guyana, and one each in Colombia, French Guyana, and
Surinam. In the Amazon watershed of Bolivia, Colombia, Ecuador, and Peru, 48 dams are
operating and 151 are planned; of these 47% were classified as “high impact” (Finer &
Jenkins, 2012; Forsberg et al., 2017). The direct and indirect damage of energy transmis­
sion lines in tropical forests is often understated (Goodland, 2005).

Oil exploration in the Amazon took off in the 1970s, mainly in Ecuador, Peru, and Bolivia
(Finer et al., 2008). A total of 263 of the 327 concessions of the Amazonia are concentrat­
ed in Andean countries. Only 25% of this total are currently in the production phase. Cur­
rent oil production activities in the Ecuador Amazon span more than 1 million hectares
and include more than 300 producing wells and 29 production camps (San Sebastián &

Page 8 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Hurtig, 2004). Practically all the Peruvian Amazon, including indigenous land but except­
ing key protected areas, is included in oil and gas exploration concessions (Finer & Orta-
Martínez, 2010). Mining in the Amazon is especially important in Brazil, where some of
the largest companies are located. There are almost 53,000 mining concessions in the
Amazon covering 163 million hectares or 21% of the Amazon—80% are in Brazil and 11%
in Peru (RAISG, 2012). A key issue is the unlawful gold digging going on almost every­
where in the Amazon: the most serious cases are in Brazil, Peru, and Guyana, with enor­
mous environmental impacts on the aquatic ecosystems, including high mercury levels in
water, soils, and humans (Dórea & Marques, 2016).

Deforestation and Forest Degradation


The road infrastructure has had a tremendous impact on the biome. Roads are the un­
equivocal main driver of deforestation (Alamgir et al., 2017; Barber, Cochrane, Souza, &
Laurance, 2014). Deforestation is obviously a direct result of agricultural expansion (Lau­
rance, Sayer, & Cassman, 2014). However, illegal gold mining deforestation is becoming
significant. The clearing process and land use is different in the Andean countries and in
Brazil and, to some extent, also in Bolivia and Colombia. In the Andean Amazon most de­
forestation has been caused for years by poor small farmers who practice different modal­
ities of shifting cultivation. After a few years of such treatment most of the cleared land is
degraded and abandoned or transformed into low productivity grassland (Shane, 1980;
Watters, 1971). African grasses cover 60% or more of the deforested Amazon. Another
portion of the land has been transformed into mid-size farms that have been established
to produce coffee and cacao or fruits; more recently, larger estates have begun planting
oil palm, among other industrial crops for export, including the illegal and highly environ­
mentally impacting coca (Araujo, 2001; Dourojeanni, 1992; Gilbert, 2012). However, in
general, properties in this region are relatively small. In Brazil and in the Bolivian low­
lands, where private ownership is more prevalent and flatland abundant, large estates
were established initially for cattle ranching; the beasts were fed on exotic grasses. Cat­
tle ranching, with numbers of beasts currently reaching some 200 million heads in Brazil
alone, continues to be the dominant use of the land (Hecht, 1985; Hecht, Norgaard, &
Possio, 1988; Margulis, 2004; Vera-Diaz, Kaufmann, & Nepstad, 2009). Part of this land is
progressively being transformed through intensive production of commodities such as
corn, cotton, or soybean. Shifting cultivation and small farms also occur in Brazil but are
not as dominant as in the Andean countries (Kalamandeen et al., 2018).

Diverse studies have measured the relation of roads to deforestation and other environ­
mental impacts, but these vary enormously with regard to the function of the road, its
quality and maintenance, soil fertility, and time elapsed since construction, among other
factors (Alves, 2001A, 2001B; Fearnside, 2007; Laurance, Goosem, & Laurance, 2009).
Measures of annual and especially of accumulated deforestation in the Amazon are heavi­
ly influenced by factors such as the purpose of the information, selected baseline year
and assumed original size of the biome, definition of forest and of national “Amazon re­
gion,” inclusion or exclusion of secondary growth forests, tree or shrub plantations, refor­
Page 9 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

estation and agroforestry fields, and, obviously, the remote sensing technologies used
(MacDicken & Tubiello, 2015). The most frequently mentioned information is collected by
the Food and Agriculture Organization (FAO) (FAO-EC, 2012). However, this information
is often contested as it relies on government information (Kim, Sexton, & Townshend,
2015). Since the beginning of the present century, deforestation information has become
more precise and independent thanks to new remote sensing technologies to address re­
duced emissions from deforestation and forest degradation (Goetz et al., 2015).

It is generally assumed that 18% to 20% of the Amazon biome has been already deprived
of its original forests (Lovejoy, 2015; Lovejoy & Nobre, 2018). Guyana, Suriname, and
French Guyana are the less deforested countries while Ecuador, with 38%, is proportion­
ally the most deforested (Izko, 2012). These statistics are driven by Brazil’s deforestation,
which covered 19.1% in 2017 (Branford & Torres, 2017), and Peru’s, with probably as
much as 17.9% (Dourojeanni, 2011A, 2018). Official and other information, using differ­
ent criteria, gives lesser proportions (Ministerio del Ambiente Peru [MINAM], 2010; Por­
tugués & Huerta, 2005; RAISG, 2015).1 It is important to mention that especially in An­
dean countries, such as Peru, only 20% to 25% of the deforested land produces crops
yearly, including grasslands, and that every hectare produces a fraction of its potential
capacity if adequately managed (Dourojeanni, 1990). This results in a gross waste of land
already deforested and serviced by roads. Also, in Andean countries most deforestation
occurs on steep hills whose soils are technically unsuitable for agriculture and by law
must be protected.

Agriculture and the keeping of livestock are the overwhelming purposes of deforestation.
But some other land uses may also be significant. It is estimated that the Brazilian hydro­
electric expansion program may occupy 9.4 million hectares (Little, 2013). Oil and gas ex­
ploitation have no significant direct impacts on deforestation, but they are indirectly im­
portant when roads are built to facilitate exploitation or to build pipelines that also favor
the invasion of loggers and farmers (Goodland, 2005). Mining, especially unlawful mining
in river banks such as in Madre de Dios, Peru, also cause direct deforestation that is in­
creasingly destructive (Asner, Llactayo, Tupayachi, & Ráez-Luna, 2013; Finer & Novoa,
2017). Urban expansion is another contributor to deforestation. However, all these land
uses, especially oil exploitation and mining, are much more significant as contributors to
forest and aquatic ecosystems’ degradation through diverse forms of contamination
(Donovan, 2012).

Information on forest degradation is obviously much less precise and difficult to define
and evaluate than deforestation (Lund, 2009; Sasaki & Putz, 2009). Selective logging, the
most common form of forest exploitation in the Amazon, is the main cause of forest degra­
dation (Foley et al., 2007; Souza & Roberts, 2005). This process occurs over decades, as
selective logging implies waves of exhaustive timber extraction that target different
species, often in the same places. Logging is complemented by hunting (Burivalova, Sek­
ercioglu, & Koh, 2014). Another key factor in degradation is the border effect that in­
creases with forest fractioning and isolation (Haddad et al., 2015), especially as related to

Page 10 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

the use of fire (Branford & Torres, 2017; Sanford, Saldarriaga, Clark, Uhl, & Herrera,
1985).

Forests are also degraded by many other exploitation modalities, like rubber tapping or
gathering of seeds and nuts and hunting, as well as by contamination originating in oil
and mining exploitation, agriculture, industry in urban areas, and even well-intentioned
ecotourism (Ceballos-Lazcurain, 1996; Peres, 2000; Peres et al., 2003; Swenson, Carte,
Domec, & Delgado, 2011). Hunting for food and commerce are today, as in the past, com­
mon activities all over the Amazon (Pierret & Dourojeanni, 1966, 1967; Nasi, Taber, & van
Vliet, 2011) and is often associated with narcotic and arms traffic (Sinovas, Price, King,
Hinsley, & Pavitt, 2017). Uncontrolled fisheries are also contributing to create “empty”
forests (Terborgh, 1999; Wilkie, Bennett, Peres, & Cunningham, 2011). It has been esti­
mated that forest degradation yearly has an impact that is equivalent to deforestation
(Carnegie Institution News, 2005). Considering the history and density of logging in Peru
it is probable that no less than half the remaining forest of this country has been degrad­
ed to some extent (Dourojeanni, 2011A).

Science and Its Influence in Amazon Conserva­


tion
The Amazon has been open to scientists from everywhere, who have worked all over the
region in every country, based in universities and research institutions from the Amazon
countries and abroad. However, they have been especially concentrated in two facilities
located in the region: the Instituto de Pesquisas da Amazônia (INPA) in Manaus, Brazil,
and the Cocha Cashu Research Station in the Manu National Park in Peru. The Instituto
Nacional de Pesquisas Espaciais (INPE) of Brazil also plays a key role with regard to cli­
mate change and vegetation monitoring, while the Instituto de Investigaciones de la Ama­
zonia Peruana (IIAP) is acquiring relevance.

Scientific advances have influenced everything that has been proposed and achieved to
conserve the Amazon. But, as will be shown in this analysis, science’s influence has been
limited and often concentrated in protected areas, mainly used to provide arguments for
their establishment and, to a lesser degree, for their management. Research has also pro­
vided inputs on small-scale projects that combine socioeconomic objectives with environ­
mental care. To a much lesser extent it has also contributed to the adoption of specific en­
vironmental legislation. Such legislation has been only partially applied and has never
been able to consistently oppose or counteract legislation that promotes business as usu­
al (Fearnside, 2014). Research has also opened new opportunities for international financ­
ing, especially as related to climate change. However, most scientific evidence was not
used to adopt policies or plans or to take action that would have real and perdurable posi­
tive impact on sustainable development (Fearnside, 1986). Instead, science and associat­
ed technology have been the main drivers of deforestation and conventional development.

Page 11 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Since the late 2000s, the most common argument to establish protected areas in the
Amazon has been the demonstration of the magnitude (“megadiversity”), uniqueness, and
patterns of distribution of their biological diversity urged by a demonstration of the grow­
ing risks of species extinction (i.e., International Union for Conservation of Nature [IUCN]
Red Lists), coupled with their much promoted actual or potential usefulness in agricul­
ture, medicine, or other areas (Myers, 1984; Mittermeier, Robles, & Mittermeier, 1997;
Reid & Miller, 1989; Schultes, 1979). Evidence of speciation and endemism in the Andes
(Swenson et al., 2012), theories such as the Pleistocene refuges (Bush, 1994; Salo, 1987),
the existence of “hot spots” (Mittermeier, Myers, Thomsen, da Fonseca, & Olivieri, 1998;
Myers, 1988), followed by the concept of cold spots (Kareiva & Marvier, 2003; Melián et
al., 2015), as well as isolation and habitat fragmentation theories (DeFries, Hansen, New­
ton, & Hansen, 2005; Haddad et al., 2015; Wilcox, 1980), minimal critical size and border
effect (Laurance et al., 2002), among others, were arguments frequently utilized and up
to some point accepted to create more and larger protected areas (Laurance et al., 2011;
Myers, 1988). These scientific discoveries also had input into the adoption of buffer zones
and the proposals of biological corridors (Beier & Noss, 2008). The intensification of field
research allowed the rediscovery of presumed extinct species and the discovery of 1,200
new species from 1999 to 2009 and 441 between 2010 and 2013 (World Wildlife Fund
[WWF], 2009; Tickell, 2013). Studies also recognize ecosystems that had been ignored for
many years, such as the Amazon white sands (Frasier, Albert, & Struwe, 2008), suggest­
ing the need for more protected areas. But the establishment of protected areas may also
be in response to opportunism or political will (Dourojeanni & Padua, 2007, 2013).

However, the scientific community has also often aired views that have provided argu­
ments to policymakers opposed to the establishment of protected areas. A pervasive con­
frontation between natural and social scientists’ views on nature and its conservation has
compromised achievements regarding protected areas. Since the late 1970s, and espe­
cially after the consecration of the concept of sustainable development (World Commis­
sion on Environment and Development [WCED], 1987), social scientists’ criticism of
strictly protected areas increased (Machlis, 1992; Machlis & Tichnell, 1985; Poole, 1989);
often they simply inverted the sense of the scientific arguments used to promote preser­
vation, such as arguing that biodiversity in protected areas is like an island’s, condemned
to extinction (Quammen, 1996). This trend has been evolving into today’s accentuated an­
thropocentrism, which, in many ways, agrees with positions that are also defended by de­
velopers (Barborak, 1997; Gómez-Pompa & Kahm, 1992; Kareiva, Marvier, & Lalasz,
2012). Another consequence of these views has been the establishment of an array of new
“protected” areas open to human occupation and a wide range of resources utilization
(Allegretti, 2008; Dudley & Stolton, 2007). On the more clearly positive side, social scien­
tists have influenced the development of so-called community-based conservation and, as
expected, they were key in the increased priority given to the recognition of rights for in­
digenous people.

Research related to forestry, freshwater vulnerability, and fisheries or wildlife (Castello et


al., 2013), and related propositions of measures to improve their management, has had
little practical acceptance. Information on natural forest regeneration (Hartshorn, 1980;
Page 12 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Whitmore, 1978) and composition (Steege et al., 2013), fisheries (Goulding, 1981; Gould­
ing, Smith, & Mahar, 1995), and wildlife (Dourojeanni, 1985B) has never been applied to
improve the management of these resources nor their legislation or administration. In­
stead, several biological discoveries were used to intensify industrial silviculture or aqua­
culture that may be detrimental to nature and natural stocks. In addition, scientific mixed
messages, such as the concept of the non-renewability of the tropical forests, were fre­
quently used as arguments to not manage forests (Gómez-Pompa, Vasquez-Yañez, & Gue­
vara, 1972).

The most transcendental scientific advances regarding the Amazon are obviously those
related to biogeochemical cycles, especially carbon fixation in the biomass and soils (Mc­
Clain, Victoria, & Richey, 2001) and the hydrological cycle (Salati & Vose, 1984; Salati,
Dall’Olio, & Matsui, 1979). These discoveries, amid others, nourished the concept of envi­
ronmental services and of their real value for world economics (Costanza et al., 1997; Far­
ber, Costanza, & Wilson, 2002) and are essential elements of the climate change issue.
For a long time, the half-truth of the Amazon as the planet’s lungs or key oxygen produc­
er dominated the scene. However, different lines of research coincided to describe a far
more complex reality. Early studies on biomass (Fitkau & Klinge, 1973; Klinge, Rodriguez,
Bruning, & Fitkau, 1975) were complemented over time by many studies on the relation
of Amazon deforestation to the global carbon problem (Baccini et al., 2012; Fearnside,
1985; Harris et al., 2012; Houghton, 2003; Houghton et al., 2000; Revelle, 1982; Wood­
well, 1978; Woodwell et al., 1983). Carbon emissions from Amazon artificial lakes and hy­
dric energy generation proved to be significant as well (Fearnside & Pueyo, 2012). Car­
bon is also accumulated in enormous quantity in Amazon soils and subsoils and implies
serious risks of emissions depending on future land use (Lahteenoja et al., 2012). Re­
search on tree physiology (Makarieva & Gorshkov, 2007; Makarieva et al., 2014), the role
of biogenic nuclei of clouds and precipitation of rain (Pöschl et al., 2010), deforestation
and fires (Nepstad et al., 2004; Koren, Kaufman, Remer, & Martins, 2004; Nepstad et al.,
1999, 2004), the so-called “flying rivers” (Newell & Newell, 1992), and extreme climatic
events (Marengo, Tomasella, Soares, Alves, & Nobre, 2011) is increasing awareness of
the potential desertification of the Amazon but also regarding a reduction of precipitation
in other South American regions (Nobre, 2014). Recent research demonstrates not only
that Amazon droughts are increasing in frequency and intensity, but also that tree com­
munities aren’t keeping up with the rapid changes. Rising carbon dioxide in the atmos­
phere is driving compositional changes in the forest (Esquivel-Muelbert et al., 2018; Lau­
rance et al., 2004). The climate change context calls for a review of protected area effi­
ciency in preserving biological diversity (Brodie, Post, & Laurance, 2012). Research has
also documented a progressive decline of the Amazon carbon sink capacity (Brienen et
al., 2015). The enormous amount of information on these matters and the growing evi­
dence of the socioeconomic consequences of deforestation are raising awareness, espe­
cially in developed countries, and channeling more resources, but have not determined
any concrete or effective decision in the Amazon countries, which continue with business
as usual, expanding the road network and indirectly promoting deforestation.

Page 13 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Before the 1960s the predominating concept regarding the Amazon’s land use capacity
was that the region had very limited potential for clean tilled agriculture and quite limit­
ed potential for permanent crops and pastures (Natural Research Council [NRC], 1982).
Most studies estimated Amazon capability for agriculture and cattle ranching at less than
11% (Zamora, 1971). This has been reflected in the legislation of countries such as Peru,
restraining deforestation. However, most soil scientists and agronomists were critical of
this view and supported the principle that Amazon soils’ natural limitations could easily
be overcome with appropriate technology that depends on economics (Sanchez & Buol,
1975; Sanchez, Bandy, Villachica, & Nicholaides, 1982). Despite justified doubts about its
sustainability (Fearnside, 1987) this trend has dominated, and large portions of the Ama­
zon, especially in Brazil, are now utilized for intensive mechanized agriculture. Advances
in agricultural sciences and technologies, including weed and pest control, soils manage­
ment, and genetically improved plants, among others, made this an economically viable
reality thanks to world food demand and new transport infrastructure. No less important
has been the enormous progress made recently regarding remote sensing, including the
use of drones, new forms of communication, building technologies, and so many others.
These may, of course, be useful to conserve the Amazon but can equally be used to deep­
en its exploitation, as in the case of geological surveys that unveil mineral richness. Other
studies have explained long-term mysteries such as the resurgences of fertility in the
northern Amazon (Bristow, Hudson-Edwards, & Chappell, 2010; Chin et al., 2015).

In conclusion, research has fulfilled its role of informing and alerting society about the
potential negative consequences of the way the Amazon is being occupied and developed,
and has proposed alternatives. The portion of the research that has been applied as sug­
gested is not negligible, especially regarding protected areas, a portion of the legislation,
and, possibly, regarding future financing of measures related to climate change. Never­
theless, society has overwhelmingly neglected scientific facts and, instead, has made in­
tense use of branches of the science and technology that contribute to business as usual
in the Amazon.

Amazon Conservation Efforts: Evolution and Si­


tuation
The careless development rush after 1945, as well as increasing deforestation and clear
evidence of social and economic failures, gave rise to multiple cautionary voices and an
array of proposals to avoid the worst expected scenarios (Dorst, 1970; Goodland & Irving,
1975). By the 1970s three strategies were given priority: the establishment of a network
of representative protected areas, the adoption of forest management as an alternative to
deforestation for agriculture, and efforts to restrict agriculture to soils that were consid­
ered suitable for such use. But by the end of the 1980s it was evident that only the first
strategy had had some success (Dourojeanni, 1990). Then, in light of new facts and politi­
cal contexts, during the 1990s the strategic choices to conserve the Amazon were diversi­
fied, covering: (a) more and larger but effectively managed protected areas, including bi­

Page 14 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

ological corridors and buffer zones; (b) the recognition of indigenous peoples’ rights over
larger territories supposed to be preserved; (c) a series of “sustainable use” options such
as “sustainable” forestry, agroforestry, and forest restoration, most of which were includ­
ed in the concept “community-based conservation”; (d) the application of land-use plan­
ning or economic-ecological zoning; (e) the financing of conservation through internation­
al technical cooperation, debt swaps, and, more recently, climate change’s related finan­
cial mechanisms; and (f) better and better enforced planning, legislation, and gover­
nance, including improved monitoring.

Protected Areas and Indigenous Land

Only 12 protected areas existed in Latin America before 1949, none of them in the Ama­
zon. The first five protected areas in the Amazon biome were established in the 1960s:
one in Bolivia, two in Peru, and two in Suriname (IUCN, 1971). But, since 1970, the estab­
lishment of larger and better-designed protected areas has been exponential, especially
in Brazil (Padua & Coimbra, 1979) and Peru (Dourojeanni, 1982; Dourojeanni & Ponce,
1978), covering around 23.5 million hectares in 1980 (3.3% of the Amazon basin) and as
much as 32.2 million hectares in 1990 (Dourojeanni, 2018; IUCN, 1990; Rojas & Castaño,
1990). By 1996, 43.8 million hectares, or 6.2% of the Amazon basin, were protected. In
2015, more than 174 million hectares (22.3% of the biome) were preserved (RAISG,
2015), of which 100 million hectares were in Brazil (RAISG, 2017; Veríssimo, Rolla, & Ve­
doveto, 2001; Verissimo & Nussbaum, 2011). A portion of this total, covering 5% of the
Amazon, pertains to transitory categories that may revert to other uses. However, ecolog­
ical representativeness of existing protected areas is still below the amount required to
ensure conservation (Fajardo, Lessmann, Bonaccorso, Devenish, & Munoz, 2014).

The ecological representativeness of the protected areas established up to 1990 was


mostly based on the theory of the Pleistocene refuges in Brazil (Wetterberg, 2004; Wetter­
berg et al., 1977) and used Holdridge’s life zones (Tosi, 1960) in the Andean countries
(Dourojeanni, 1968, 1976, 2018; Dourojeanni & Ponce, 1978). More sophisticated meth­
ods were later used in Brazil, such as gap analysis in large meetings of specialists (Min­
istério do Meio Ambiente [MMA], 2001) and, more recently, applying the methodology of
rapid biological assessment (Sayre et al., 1999), a method that, since the late 1990s, has
been widely used in Andean Amazon countries. Nonetheless, protected areas’ representa­
tiveness and defensibility are still below requirements (Coetzee, Gaston, & Chown, 2014;
DeFries et al., 2005; Laurance et al., 2012; Peres & Terborgh, 1995). Until 1990, well over
90% of the protected areas pertained to categories I to IV of IUCN, meaning that no peo­
ple or utilization of resources were allowed inside. But 46% of the protected areas now
pertain to categories that allow the presence of humans and resources exploitation. This
has been essentially a result of: (a) the development of a strong anthropocentric social
environmentalism, especially but not only in Brazil, that promotes the concept that tradi­
tional people better conserve nature than protection designations (Daniels, 2002;
Diegues, 2005; Fairhead, Leach, & Scoones, 2011; White, Khare, & Motnar, 2005) and (b)
the growing limitation of areas free of human occupation or development interests to es­
tablish conventional protected areas. Some categories of protected areas, such as the
Page 15 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Brazilian extractive reserves (Allegretti, 2008), where logging and farming are allowed,
have, as their primary objective, to allow traditional peoples access to land and re­
sources, while others, such the large Áreas de Proteção Ambiental, are open to almost
every possible legal use, including agriculture, villages, and rural industries. Additionally,
this group includes national forests that are leased to logging concessions. Therefore, de­
spite numbers being impressive, a clearly insufficient 11.8% of the Amazon is represent­
ed in fully protected areas.

Management effectiveness of Amazon protected areas of all categories is limited due to


low political priority, corresponding funding shortages, and lack of staff, infrastructure,
and equipment (Dourojeanni, 2002; Dourojeanni & Quiroga, 2006). Management plans,
when available, are not economically feasible, and are often outdated and rarely applied.
Many parks are quite isolated and not yet open to visitation, so are not contributing to lo­
cal development despite strong evidence of the economic advantages of the environmen­
tal services they provide and their tourism potential (León, 2007; Young & Medeiros,
2018). All categories of protected areas are increasingly threatened by the surrounding
population, poverty, and development needs, such as roads and dams, which are responsi­
ble for more frequent decisions of downgrading, downsizing, and degazettement annually,
especially in Brazil (Bernard, Penna, & Araujo, 2014; Fearnside & Vilela, 2018; Mascia et
al., 2014). Despite divergent views that support opening strictly protected areas for sus­
tainable uses by traditional populations (Chapin, 2004; Dowie, 2005; Galvin & Haller,
2008; Mora & Sale, 2011; Porter-Bolland et al., 2012), the evidence is that national parks
and other strictly protected designations have been and remain the most effective and ef­
ficient options to conserve viable samples of the ecosystems and the biodiversity they
contain (Andam, Ferraro, Pfaff, Sanchez, & Robalino, 2008; Bruner, Gullison, Rice, & da
Fonseca, 2001; Chape, Harrison, Spalding, & Lysenko, 2005; Dourojeanni, 2015; Lau­
rance et al., 2012; Nepstad et al., 2006). Instead, evidence is being accumulated regard­
ing the inherent difficulty of protecting some biological resources while using others in
the same area (Nepstad, 1997; Peres, 2000; Redford & Stearman, 1993). It is also clear
that human occupation of both types of protected areas is necessary and may be comple­
mentary (Dourojeanni & Padua, 2007; Peres, 2011).

As result of the evidence on the impact of isolation and lack of connectivity aggravated by
the risks of global warming (Damschen, Haddad, Orrock, Tewksbury, & Levey, 2006) ef­
forts were made to establish transboundary protected areas (International Tropical Tim­
ber Organization-United Nations University [ITTO-UNU], 2011), parks for peace (Ali et
al., 2006), and, especially, biological corridors (Bennet, 2005; Laurance, Vasconcelos, &
Lovejoy, 2001; Rosenberg, Noon, & Meslow, 1997) in the Amazon. Several biological corri­
dors were proposed between large protected areas in every country and among those in
neighboring countries. However, these projects have not been officially endorsed and face
enormous practical obstacles and scientific challenges (Haddad et al., 2014). A significant
contribution to Amazon conservation is the growing number of privately owned protected
areas as well as the Peruvian forest concessions designed for private conservation or eco­

Page 16 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

tourism operations. The long-term fate of protected areas is unclear (Laurance et al.,
2012).

Brazil has been pioneering the recognition of indigenous territories since 1910. However,
until the 1970s the Amazon was viewed as an empty space (Smith, 1983). While Brazil al­
ready had some large indigenous territories (mostly reserves), other countries considered
native people as negligible or as simple farmers and provided them with only very small
patches of land. In the 1980s, the Amazon indigenous people, with a population estimated
at 2 million to 3 million, many of whom live in urban areas, spoke out strongly and began
to play an increasingly important political role in almost every country, with strong inter­
national support. In 2015, recognized indigenous land amounted to 168.4 million
hectares, while 5.1 million hectares more were to be recognized or pertained to other cat­
egories, totaling 28.1% of the Amazon biome. Brazil alone had 118 million hectares main­
ly under the category of indigenous reserve. Venezuela, Colombia, Peru, and Bolivia fol­
lowed in indigenous land size. Indigenous land represents 72% of the Venezuelan Amazon
and above 50% of the Colombian and Ecuadorian Amazon (RAISG, 2015). Part of these
territories is superposed on protected areas. Indigenous claims for more land, including
inside protected areas, continues in every country.

With regard to native peoples’ rights over the land, a strong argument used to expand
their territories is the value of their assumed traditional knowledge to wisely manage nat­
ural resources (Plotkin, 1993). Their lands are still covered by forests and, in many cases,
they have efficiently resisted the advances of agriculture and logging. However, they re­
veal growing susceptibility to the temptation of economic benefits through conventional
development, including rental of the land for agriculture and livestock, logging, and min­
ing. There is an ongoing debate on the long-term validity of these lands with regard to
conserving nature (Alcorn, 2010; Clay, Alcorn, & Butler, 2000), but it is unquestionable
that they currently offer a very important opportunity for conservation provided titling,
technical assistance, and other measures are opportunely put in place (Blackman, Corral,
Lima, & Asner, 2017; Daniels, 2002; Dourojeanni, 2011A; Garnett et al., 2018).

Considering protected areas and indigenous territories, but excluding their reciprocal su­
perposition, no less than 45.4% of the Amazon is somehow protected (RAISG, 2017). De­
spite all mentioned limitations and growing challenges, the establishment of a system of
protected areas complemented by recognized indigenous land is the clearest, largest, and
probably most perdurable result of past conservation initiatives. Efforts must be made to
expand strictly protected areas and to improve conservation through effective resources
management within indigenous territories and protected areas that are legally open to
exploitation.

Sustainable Use Options

While it is evident that protected areas in the Amazon are an essential but insufficient
component of any strategy to limit deforestation and degradation (Soares-Filho et al.,

Page 17 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

2006), several other options are being applied to the end goal of achieving sustainable de­
velopment.

Forest management for sustainable timber production has been the first choice to avoid
deforestation in the Amazon. This option has been promoted by the Food and Agriculture
Organization (FAO) since the 1950s. National forests and reserves were established even
before protected areas, and several important forest management projects were conduct­
ed in Brazil, Peru, and Venezuela, especially in the 1970s (Dourojeanni, 1999; Lentini,
Pereira, Celentano, & Pereira, 2005; Schmidt, 1987). As experiments, they demonstrated
that tropical forest management is technically viable, economically profitable (Alavapari,
Putz, & Schmink, 2004), and could be environmentally sound (Fearnside, 2008). However,
they all failed, as the national social and related economic contexts made it impractical to
sustainably produce timber while unlawful chaotic logging, often associated with un­
planned deforestation, was going on everywhere. Social disorder made it almost impossi­
ble to avoid the landless farmer’s invasion of legal long-term forest concessions (Douro­
jeanni, 2009B). Selective logging, provided technically determined rotations are respect­
ed, may preserve environmental services, but this is rarely the case (Asner et al., 2005;
Edwards, Tobias, Sheil, Meijaard, & Laurance, 2014; Nepstad et al., 1999; Putzel, Peters,
& Romo, 2011). Since the 1960s forest laws have been regularly modernized in every
country, sustainability criteria and indicators adopted (Pokorny, Sabogal, Prabhu, & Silva,
2002), forest certification promoted (Agrawal, Chhatre, & Hardin, 2008), and some low-
impact logging has been practiced (Putz, Sist, Fredericksen, & Dykstra, 2008). However,
as of 2005, even the most optimistic evaluation found that only 3.5% of the production
forests of Latin America and the Caribbean were managed to an extent and only 60% of
this fraction had forest certification (International Tropical Timber Organization [ITTO],
2006). Corruption in the Amazon forestry sector is increasing (Finer, Jenkins, Blue Sky, &
Pine, 2014; Urrunaga, Johnson, Orbegozo, & Mulligan, 2012) and new agronomical tech­
niques and opportunities are seriously challenging natural forestry’s competitive prof­
itability if its environmental services are not taken into account. As of today, by every pa­
rameter sustainable management of natural forests in the Amazon has failed. Instead, sil­
viculture with exotic or native species is progressing.

Agroforestry, the combination of trees and crops in the same land, has been considered
an alternative use of the land that is less environmentally destructive than conventional
agriculture or cattle ranching. This has considerably expanded, assuming, especially in
the Andean countries, the form of understory coffee and cacao plantations (Miller & Nair,
2006; Somarriba et al., 2012). Many other agroforestry practices exist that combine na­
tive or exotic trees with crops (Dubois, Viana, & Anderson, 1996; Pinho, Miller, & Alfaia,
2012). While it is evident that agroforestry is environmentally friendlier than open field
agriculture, especially regarding soils and water resources, its environmental benefits do
not match that of the original forest (Fearnside, 1995). Additionally, some commonly pro­
moted definitions of agroforestry are equivalent to shifting cultivation, implying land
clearing (Dourojeanni, 2009A).

Page 18 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Large portions of the Amazon are covered by secondary forests that usually regenerate
well (Dent & Right, 2009; Sanchez-Cuervo, Aide, Clark, & Etter, 2012) if use of the land
has not been excessively intense (Jakovac, Pena-Claros, Kuyper, & Bongers, 2015). The
permanence or age of the fallow forests decreases as land use intensifies (Carreiras,
Pereira, Campagnolo, & Shimabukuro, 2006); nevertheless, these forests are significant
as carbon sinks and to regulate water flux, among other environmental services. It has
been estimated that 13.2 million hectares of these forests existed in Brazil in 2006 (Almei­
da, Valeriano, Escada, & Rennó, 2010) and proportionally much more in Peru and other
Andean countries. The management of fallow forests or secondary growth to produce
fast-growing timber (Dourojeanni, 1987; Emrich, Pokorny, & Sepp, 2000) or other goods
is an interesting option that is beginning to be applied (Chokkalingan & de Jong, 2001;
Kammesheidt, Köhler, & Huth, 2002). Despite being heavily anthropic, these forests con­
serve a significant portion of their original biodiversity (Almeida et al., 2010; Chazdon et
al., 2009; Dent & Right, 2009; Peres et al., 2010), but they do not replace all services pro­
vided by natural forests (Gibson et al., 2011; Jakovac et al., 2015).

It is often argued that industrial forest plantations and tree crops, including oil palm, are
a better use of the Amazon land than clean tilled agriculture or pasture. Permanent plan­
tations are environmentally better than annual crops and theoretically they may spare
forests (Gutierrez-Velez et al., 2011). But this is not an accepted argument when these
ventures require deforestation of natural stands of trees (Dammert, 2013). Of course,
there are many ways of establishing plantations and some are less environmentally dam­
aging than others. Forest restoration or rehabilitation in the Amazon is incipient and used
on a minimal scale, especially in protected areas or as obligations imposed by environ­
mental conditionality. Evidence exists that natural regeneration is the cheapest and most
effective option to restore vegetation, if cattle and fire are excluded (Sanchez-Cuervo et
al., 2012; Zahawi, Holl, Cole, & Reid, 2013).

Diverse locally adapted combinations of the previous options, often mixed with fisheries
and wildlife management and ecotourism, especially inside and nearby protected areas,
have been applied in the Amazon since the late 1980s. They are part of a wave of projects
that are known as “community-based conservation” when used by traditional or indige­
nous peoples, who were, frequently, the originators of such concepts. This strategy is
based on the theory that conservation and development could be simultaneously
achieved, but for ideological reasons it has been widely used to promote the direct use of
the protected areas’ resources, even their extinction (Ghimire & Pimbert, 1997; Mur­
phree, 2002; Santilli, 2005). Despite social scientists attributing a wide range of environ­
mental and socioeconomic virtues to community-based conservation these are not so obvi­
ous when scrutinized and many difficulties and shortcomings become evident (Agrawal &
Gibson, 1999; Barrett, Brandon, Gibson, & Gjertsen, 2001; Brooks, Waylen, & Borgerhoff
Mulde, 2013; Dourojeanni, 2008; Kellert, Mehta, Ebbin, & Lichenfeld, 2000; Redford &
Padoch, 1992; Schwartzman, Nepstad, & Moreira, 2000). But the community-based con­
servation concept is not only related merely to protected areas; it is used and may be use­
ful over a wider space (Berkes, 2004). It is seen as an especially promising option for

Page 19 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

communities in indigenous territories and in buffer zones around protected areas (Douro­
jeanni, 2011A).

It is difficult to precisely calculate the extent of the Amazon that is covered by each of the
several types of land use, and especially by those that may be considered sustainable, but
the evidence is that forest losses are prevalent despite some gains (Hansen et al., 2013).
There is no match among the agricultural census, information on deforestation, and sta­
tistics on protected areas and indigenous lands, nor with information on secondary
forests. Also, each country uses different approaches. But by any measurement it is evi­
dent that the Amazon is under very heavy pressure and that at the national level most
Amazon countries already have proportionately less forested area—including reconstitut­
ed—than developed countries such as Japan or Sweden. Furthermore, in most developed
countries the forest area is increasing (Veríssimo & Nussbaum, 2011).

Policy, Planning, Legislation, Enforcement

Another set of strategies applied to conserve the Amazon is, of course, a combination of
better-adapted policies, long-term planning, adequate legislation, and enforcement using
well-known tools such as institutional strengthening and governance improvement with
more and better information and participation, monitoring, and control. Important efforts
were developed in almost every Amazon country, sometimes under pressure and with the
support of developed countries to improve sustainable development and conservation in
the region. Much progress has been achieved: more comprehensive legislation, especially
regarding environmental licensing of infrastructure, rules such as the percentage of land
that must be kept under natural forest in private properties, the compulsory protection of
river banks and other erodible terrains, the establishment of protected areas, and the
growing recognition of indigenous rights are a few among many other results. Brazil is
the country where progress in these aspects has been most noticeable. Despite legisla­
tion prepared in capital cities often not being well adapted to Amazon reality, in general
terms it is adequate to—in theory—ensure the sustainable use of the Amazon. However,
legislation related to natural renewable resources and environment is scarcely enforced
and often not even applied (Dourojeanni, 2013B; Morim Novaes & de França Souza, 2013;
Soares-Filho et al., 2014). Institutions that directly relate to natural resources such as
forestry or fisheries have insufficient and irregular budgets and chronic lack of staff,
while those who work in these industries are usually underpaid and susceptible to corrup­
tion.

If legislation can broadly be considered appropriate, development policy design for the
Amazon is still undefined between the majority interested in business as usual and a very
small minority willing to experiment with a more sustainable style of development. Usual­
ly productive sectors such as energy and mining, as well as those in charge of infrastruc­
ture, fully dominate the government and disregard weak environmental or planning insti­
tutions (Enrique & Cueto, 2011). Policies rarely guide the planning, which is not coordi­
nated among the nation and its regions and is often clearly contradictory even among
public sectors. Additionally, planning is frequently modified to satisfy the specific and pre­

Page 20 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

dicted interests of successive governments. Since the 1980s a series of efforts to apply di­
verse modalities of land use or territorial planning, often locally known as ecological-eco­
nomical zoning, has been developed in Brazil and other countries (Acselrad, 2001). How­
ever, their fate has been the same as for policies and plans. Their benefits have been re­
stricted to a collateral improvement of environmental consciousness.

A positive development since the late 1990s has been the growing demands of local popu­
lations, especially indigenous peoples, to be informed and especially to have a say in the
decisions on public works and natural resources exploitation initiatives that affect them.
In response, most countries approved new legislation (Garcia, 2014) regarding trans­
parency and participation in public decisions and accepted the rule of the International
Labour Organization’s Decision 169 concerning indigenous and tribal rights. Indigenous
people have been politically very active since the turn of the 21st century and dispropor­
tionally influential in regional decisions considering their very small minority. Several
nongovernmental organizations are providing support to indigenous activism. However,
there is still a long way to go to achieve an equitable participation of all involved.

In conclusion, the scenario for Amazon conservation is one of very poor governance ag­
gravated by a traditionally corrupt authoritarian government style and social disorder
that does not facilitate the application of a sustainable option of development.

Financing Conservation

Resources were always discretely available to promote conservation in the Amazon. Mod­
est governmental contributions, initially through the forest services, were often matched
by international or multilateral cooperation (Leal-Farias, 2018; Mittermeier & Bowles,
1993). International nongovernmental organizations, directly or through local correspon­
dents in a changing context, also assisted in the establishment and management of many
important protected areas of the Amazon (Dourojeanni, 2006). Ideas such as the debt
swaps for nature (Post, 1990; Reilly, 2006; Visser & Mendoza, 1994) raised expectations
but, except in Peru, they were not significant. Since the 1980s the funding for protected
areas has been increasingly diverted to social issues, including “community-based conser­
vation.” The growing consensus on the important role of the Amazon forest regarding
global climate change preconized a new approach that, based on scientific but changing
evidence, proposed a payment for carbon fixation in the biomass. The idea was trans­
formed by the Kyoto Protocol (1997) and later by the Paris Agreement (2015) in a series
of tools, including the “Clean Development Mechanism,” a market for emissions, and the
concept of joint implementation. This concept has evolved to its present form (Reducing
Emissions from Deforestation and Forest Degradation or REDD) aiming at creating a fi­
nancial value for the carbon stored in forests, offering incentives for developing countries
to reduce emissions from forested lands and invest in low-carbon paths to sustainable de­
velopment. REDD+ goes beyond deforestation and forest degradation to include the role
of conservation, sustainable management of forests, and enhancement of forest carbon
stocks (Olander, Gibbs, Steininger, Swenson, & Murray, 2008; Putz & Romero, 2012). Un­
til now, despite all efforts made, this evolving initiative has not resulted in significant ad­

Page 21 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

vantages for conservation in the Amazon and, instead, has raised suspicions, especially
among indigenous people.

International funding for conservation in the Amazon comes primarily from 24 major fun­
ders (Castro de la Mata & Riega-Campos, 2014). This funding totaled approximately
US$206.2 million per year between 2007 and 2013, a fourfold increase since the 1990s.
Fifty percent of this amount goes to Brazil, followed by Peru, Guyana, Colombia, and
Ecuador. The top three organizations funding conservation in the Amazon basin are the
Norwegian Agency for Development Cooperation, the World Bank, and the Gordon and
Betty Moore Foundation. Seven of the top 10 funders are bilateral and multilateral agen­
cies. Funding is split among legislation, policy, planning, enforcement, and so on (38%);
payment for environmental services and REDD (24%); protected areas (16%); indigenous
people, local livelihood, education, and so on (14%).

It is difficult to define the share of national and subnational governments in Amazon con­
servation investments, but a review of the annual budgets of the concerned ministries
and agencies reveals they invest much more than external assistance, even when consid­
ering subjects such as protected areas that receive some international priority (Douro­
jeanni & Quiroga, 2006; Riva, da Fonseca, & Hasenlever, 2008). But foreign funds are es­
sential to cover gaps and avoid bureaucratic procedures that plague Amazonian coun­
tries’ institutions.

Some original ideas have been developed to facilitate funding of conservation projects.
Brazil’s legislation requires significant environmental compensation for infrastructure as
part of licensing, including a percent for protected areas. Also, some states are compen­
sating municipalities that have managed protected areas in their territory well with a fa­
vorable tax return. In several countries special funds have been established to finance
such environmental projects as the Fundo Brasileiro para a Biodiversidade (FUNBIO) in
Brazil or the Peruvian Trust Fund for National Parks and Protected Areas (PROFO­
NANPE) in Peru.

Financing mechanisms for Amazon conservation, both from Amazon countries and from
other nations, despite being clearly insufficient, have played a key role in what has been
achieved. The subsidies for activities that contribute to deforestation alone are much
more significant than investments in conservation, however (McFarland, Whitley, &
Kissinger, 2015). Amazon countries as well as foreigners are investing several hundred
times more in what can be called “environmentally risky developments,” such as trans­
portation (Dourojeanni, 2011A; Little, 2013), energy, mining, agriculture, and forestry ex­
pansion, than in sustainable development and conservation alternatives (Dourojeanni,
2013B; Dourojeanni, Barandiaran, & Dourojeanni, 2010; GREFI 2014).

Page 22 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

The New Amazon


Despite some progress toward sustainability there is no evidence that current develop­
ment trends in the Amazon will change in the next decade or so. All projections for the re­
gion as a whole (Freire et al., 2007; Little, 2013) as well as for every individual country or
region (Arenas et al., 2011; Dourojeanni, 2013B; Dourojeanni, Barandiarán, & Dourojean­
ni, 2010; Izko, 2012) demonstrate the existence and continuous preparation of new poli­
cies, plans, and financial availability for an enormous expansion of transportation and en­
ergy infrastructure as well as to increase export agriculture, hydrocarbon, and mining ex­
ploitation. Every prospect foresees a substantial growth of the resident population, espe­
cially urban. All scenarios, including those that are environmentally optimistic, demon­
strate increasing deforestation and forest degradation and aggravation of environmental
issues such as carbon emissions, extreme weather, alteration of the rain pattern, and bio­
diversity losses. The strong and growing concentration of the population in urban areas is
often seen as a relief for nature but the opposite may be true (Uriartea et al., 2012). Even
the past gains regarding Amazon conservation, such as protected areas and indigenous
land, are threatened by isolation and climate change but especially by social and econom­
ic pressures from inside each country, especially Brazil, and increasingly from extra-re­
gional areas such as China and other Asian economic powers, in addition to North Ameri­
ca and Europe. The optimism of experts about Amazon forest conservation is very low
(Terborgh, 1999, 2000).

It is improbable that current national or international efforts to channel more resources


to avoid the worst scenarios, including those related to climate change, will modify this
prognosis even in the unlikely case they be fully successful. Therefore, the enormous gap
between proposals of zero deforestation (Brown & Zarin, 2013; Gibbs et al., 2015) and re­
ality will continue and increase. Some foresters expect tropical forests to be reduced to a
minimal level and, instead, be substituted by a mix of anthropic ecosystems (Blaser &
Gregersen, 2013). Nature will adapt and simply be different, and some believe it will still
be promissory (Lugo, 2013).

Sustainable development in the Amazon is not a utopia (Nepstad et al., 2009). It is possi­
ble that there is enough knowledge accumulated to increase agricultural productivity sev­
eral times, to produce more energy and minerals, and to slow the exploitation of renew­
able natural resources while conserving remaining natural forests’ values and services.
Small-scale evidence of compatibility between development and environment exists
everywhere in the region and in almost every economic field. Most such schemes are
ephemeral, mainly as the consequence of an economic and social environment related to
gross disobedience of the law, but some have developed economic strategies that allow
them to endure enough to prove their validity. As shown throughout this review, the cen­
tral cause of the pessimism about achieving sustainable development in the Amazon is
the poor governance capacity and the deeply rooted social disorder that still prevail in
the countries that are responsible for the region. If public policies and plans or legislation
are not legitimized by transparency and participation and are not enforced, as today,
hope is limited. These countries are progressing in the right direction but the anarchy in
Page 23 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

the use of the resources is still not under control. Thus, the general situation will still de­
teriorate before starting to improve. But every gain for sustainability achieved today will
facilitate future recovery. Therefore, it is worth continuing to work harder.

Further Reading
Blaser, J., & Gregersen, H. (2013). Forests in the next 300 years. Unasylva, 240(64), 61–
73.

Fearnside, P. M. (1986). Human carrying capacity of the Brazilian rainforest. New York,
NY: Columbia University Press.

Fearnside, P. M. (Ed.). (2009). A floresta Amazônica nas mudanças globais. Manaus,


Brazil: Instituto Nacional de Pesquisas da Amazônia-INPA.

Goulding, M., Smith, N. J. H., & Mahar, D. J. (1996). Floods of fortune: Ecology and econo­
my along the Amazon. New York, NY: Columbia University Press.

Harcourt, C. S., & Sayer, J. A. (Eds.). (1995). The conservation atlas of tropical forests:
The Americas. New York, NY: Simon & Schuster.

Laurance, W. F., & Peres, C. A. (Eds.). (2006). Emerging threats to tropical forests. Chica­
go, IL: University of Chicago Press.

Terborgh, J. (2004). Requiem for nature. Washington, DC: Island Press.

References
Acselrad, H. (2001). O zoneamento ecológico-econômico da Amazônia e o panoptismo im­
perfeito. Planejamento e Território, 15–16(2), 53–75.

Agrawal, A., Chhatre, A., & Hardin, R. (2008). Changing governance of the world’s
forests. Science, 320, 1460–1463.

Agrawal, A., & Gibson, C. C. (1999). Enchantment and disenchantment: The role of com­
munity in natural resource conservation. World Development, 27, 629–649.

Ahmed, S. E., Souza, C. M., Jr., Ribeiro, J., & Ewers, R. M. (2013). Temporal patterns of
road network development in the Brazilian Amazon. Regional Environmental
Change, 13(5), 927–937.

Alamgir, M., Campbell, M. J., Sloan, S., Goosem, M., Clements, G. R., Mahmoud, M. I., &
Laurance, W. F. (2017). Economic, Socio-Political and Environmental Risks of Road
Development in the Tropics. Current Biology 27(20), R1130–R1140.

Alavapari, J. R. R., Putz, F. E., & Schmink, M. (Eds.). (2004). Working forests in the
neotropics. New York, NY: Columbia University Press.

Page 24 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Alberti, J., & Pereyra, A. (2018). Carretera Interoceánica IIRSA Sur de Perú: Un
megaproyecto con preinversión express. Estudios de Caso de Megaproyectos, Monografía
Banco Interamericano de Desarrollo, Washington, DC.

Alcorn, J. (2010). Indigenous peoples and conservation: A White Paper prepared for
the MacArthur Foundation. MacArthur Foundation Conservation White Paper Series
2010.

Ali, S., Sandwith, T., Besançon, Ch., Alcalde, M., Ponce, C. F., Curonisy, Y., . . . Duffy, R.
(2006). Parks for peace or peace for parks? Issues in practice and policy. ECSP.

Allegretti, M. (2008). The social construction of public policies: Chico Mendes and the
seringueiros movement. Editora UFPR, Desenvolvimento e Meio Ambiente, 18, 39–59.

Almeida, C. A., Valeriano, D. M., Escada, M. L. S., & Rennó, C. D. (2010). Estimation of
secondary vegetation area in the Brazilian Legal Amazon. Acta Amazonica, 40(2),
289–302.

Alves, D. S. (2001a). An analysis of the geographical patterns of deforestation in the


Brazilian Amazon during the 1991–1996 period. In C. Wood and R. Porro (Eds.), Patterns
and process of land use and forest changes in the Amazon (pp. 95–106). Gainesville: Uni­
versity of Florida Press.

Alves, D. S. (2001b). O processo de desflorestamento na Amazônia. Instituto de Pesquisas


Espaciais (INPE), Parcerias Estratégicas, 12, 259–275.

Andam, K., Ferraro, P. J., Pfaff, A., Sanchez, G. A., & Robalino, J. A. (2008). Measuring ef­
fectiveness of protected areas networks in reducing deforestation. PNAS, 106, 16089–
16094.

Araujo, R. (2001). Tráfico de drogas, economías ilícitas y sociedad en la Amazonia occi­


dental. UNESCO Revista Internacional de Ciencias Sociales No. 169, Paris.

Arenas, W., Zúñiga, P., & Mayordomo, E. (2011). Retos para un desarrollo sostenible:
Transformaciones en la Amazonia colombiana. Estudio de la Amazonía Colombiana. Bo­
gotá, Colombia: Fundación Alisos.

Asner, G. P., Knapp, D. E., Broadbent, E., Oliviera, P., Keller, M., & Silva, J. (2005). Selec­
tive logging in the Brazilian Amazon. Science, 310(5747), 480–482.

Asner, G., Llactayo, W., Tupayachi, R., & Ráez-Luna, E. (2013). Elevated rates of gold
mining in the Amazon revealed through high-resolution monitoring. PNAS,
September 30.

Atwood, R. (2017). Connecting Two Realms. Archaeologists rethink the early civi­
lizations of the Amazon. Archeology, July/August.

Page 25 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Atwood, R. (2011). Early pyramids: Jaen, Peru. Archeology Magazine Archives, Archae­
ological Institute of America, 64(1), 1–4.

Baccini, A., Goetz, S. J., Walker, W. S., Laporte, N. T., Sun, M., Sulla-Menashe, D., . . .
Houghton, R. A. (2012). Estimated carbon dioxide emissions from tropical defor­
estation improved by carbon-density maps. Nature Climate Changes Letters, 2, 182–
185.

Barber, P. Ch., Cochrane, M. A., Souza, C. M., Jr., & Laurance, W. F. (2014). Roads, defor­
estation, and the mitigating effect of protected areas in the Amazon. Biological Conserva­
tion, 177, 203–209.

Barborak, J. (1997). Mitos e realidades da concepção atual de áreas protegidas na Améri­


ca Latina. Anais I Congresso Brasileiro de Unidades de Conservação Curitiba, 1, 39–47.

Barni, P. E., Fearnside, Ph. M., & Graça, P. M. L. A. (2015). Simulating deforestation Ama­
zonia: Impacts in Brazil’s Roraima state from reconstructing Highway BR-319 (Manaus-
Porto Velho). Environmental Management, 55(2), 259–278.

Barrett, C. B., Brandon, K., Gibson, C., & Gjertsen, H. (2001). Conserving tropical biodi­
versity amid weak institutions. BioScience, 51, 497–502.

Basso, L. F. C., & Kimura, H. (2008). Ecossistema terra preta: Recriar o que já foi criado a
terra preta dos indios da Amazônia. Revista Ciências Administrativas, 14(2), 230–250.

Beier, P., & Noss, R. F. (2008). Do habitat corridors provide connectivity?Conserva­


tion Biology, 12(6), 1241–1252.

Belaunde, F. (1969). La conquista del Perú por los peruanos. Lima, Peru: Ed. Tahuan­
tinsuyo.

Bennet, A. F. (2005). Linkages in the landscape: The role of corridors and connectivity in
wildlife conservation. Gland, Switzerland: IUCN.

Berkes, F. (2004). Rethinking community-based conservation. Conservation Biology,


18(3), 621–630.

Bernard, E., Penna, L. A. O., & Araujo, E. (2014). Downgrading, downsizing, degazette­
ment, and reclassification of protected areas in Brazil. Conservation Biology, 28(4), 939–
950.

Blackman, A., Corral, L., Lima, E. S., & Asner, G. P. (2017). Titling indigenous commu­
nities protects forests in the Peruvian Amazon. PNAS, 114(16), 4123–4128.

Blaser, J., & Gregersen, H. (2013). Forests in the next 300 years. Unasylva, 240(64), 61–
73.

Page 26 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Blatrix, R., Roux, B., Béarez, P., Prestes-Carneiro, G., Amaya, M., Aramayo, J. L., . . . McK­
ey, D. (2018). The unique functioning of a pre-Columbian Amazonian floodplain
fishery. Scientific Reports, 8, 5998.

Brack, A. (2003). Perú: Diez mil años de domesticación. Lima, Peru: Bruno.

Brackelaire, V. (2006). Situación de los últimos pueblos indígenas aislados den América
Latina (Bolivia, Brasil, Colombia, Ecuador, Paraguay, Perú, Venezuela). In V. Brackelaire
(Ed.), Diagnóstico regional para facilitar estrategias de protección. Brazil: DF.

Branford, S., & Torres, M. (2017). Record Amazon fires stun scientists; sign of sick,
degraded forests. Mongabay, October 11.

Brienen, R. J. W., Phillips, O. L., Feldpausch, T. R., Gloor, E., Baker, T. R., Lloyd, J., . . . Za­
gt, R. J. (2015). Long-term decline of the Amazon carbon sink. Nature, 519(7543),
344–348.

Bristow, C. S., Hudson-Edwards, K. A., & Chappell, A. (2010). Fertilizing the Amazon
and equatorial Atlantic with West African dust. Geophysical Research Letters, 37,
L14807.

Brodie, J., Post, E., & Laurance, W. F. (2012). Climate change and tropical biodiversity: A
new focus. Trends in Ecology and Evolution, 27(3), 145–150.

Brooks, J., Waylen, K. A., & Borgerhoff Mulde, M. (2013). Assessing community-based
conservation projects: A systematic review and multilevel analysis of attitudinal,
behavioral, ecological, and economic outcomes. Environmental Evidence, 2, 2.

Brown, S., & Zarin, D. (2013). What does zero deforestation mean? Science, Environmen­
tal Science, 342, 806–807.

Bruner, A. G., Gullison, R., Rice, E., & da Fonseca, G. A. B. (2001). Effectiveness of parks
in protecting tropical biodiversity. Science, 291(5501), 125–128.

Bunker, Stephen G. (1984). Modes of extraction, unequal exchange, and the progressive
underdevelopment of an extreme periphery: The Brazilian Amazon, 1600-1980. American
Journal of Sociology, 89(5), 1017–1064.

Burga, M. (2008). Choquequirao; símbolo de la resistencia andina (historia, antropología


y linguística). Lima, Peru: Universidad Nacional Mayor de San Marcos.

Burivalova, Z., Sekercioglu, C. H., & Koh, L. P. (2014). Thresholds of logging intensity
to maintain tropical forest biodiversity. Current Biology, 24(16), 1893–1898.

Bush, M. B. (1994). Amazonian speciation: A necessarily complex model. Journal of Bio­


geography, 1(21), 5–17.

Butler, R. (2017). Calculating deforestation figures for the Amazon. Mongabay, Janu­
ary 26.
Page 27 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Caillaux, J., Novak, F., & Ruiz, M. (2016). Las relaciones de China con América Latina y el
ferrocarril Bioceánico Brasil-Perú. Lima, Peru: Sociedad Peruana de Derecho Ambiental e
Instituto de Estudios Internacionales.

Carnegie Institution News. (2005). Selective logging disturbance in the Brazilian Amazon:
New satellite study doubles forest disturbance estimates in Brazil—impacts widespread.
Carnegie Institution News, October 20.

Carreiras, J. M. B., Pereira, J. M. C., Campagnolo, M. L., & Shimabukuro, Y. E. (2006). As­
sessing the extent of agriculture/pasture and secondary succession forest in the Brazilian
Legal Amazon using spot vegetation data. Remote Sensing of Environment, 101(3), 283–
298.

Castello, L, McGrath, D. G., Hess, L. L., Coe, M. T., Lefebvre, P. A., Petry, P., . . . Arantes,
C. C. (2013). The vulnerability of Amazon freshwater ecosystems. Conservation Let­
ters, 1–13.

Castro de La Mata, G., & Riega-Campos, S. (2014). An analysis of international conserva­


tion funding in the Amazon. Washington, DC: Ecosystem Services.

Ceballos-Lazcurain, H. (1996). Tourism, ecotourism and protected areas. Gland, Switzer­


land: IUCN.

Chape, S., Harrison, J., Spalding, M., & Lysenko, I. (2005). Measuring the extent and ef­
fectiveness of protected areas as an indicator for meeting global biodiversity targets.
Philosophical Transactions of the Royal Society of London, B360, 443–455.

Chapin, M. (2004). A challenge to conservationists: Can we protect natural habitats with­


out abusing the people who live in them? World Watch Magazine, 17(6), 17–31.

Chazdon, R. L., Letcher, S. G., van Breugel, M., Martinez-Ramos, M., Bongers, F., Lamb,
D., . . . Miller, S. E. (2009). The potential for species conservation in tropical secondary
forests. Conservation Biology, 23, 1406–1417.

Chin, M., Ginoux, P., Kinne, S., Hongbin Yu, O., Chin, M., Yuan, T., . . . Zhao, C. (2015). The
fertilizing role of African dust in the amazon rainforest: A first multiyear assessment
based on Calipso lidar observations. Geophysical Research Letters, 42(6), 1984–1991.

Chokkalingam, U., & de Jong, W. (2001). Secondary forest: A working definition and typol­
ogy. International Forestry Review, 3(1), 19–26.

Church, W. B. (1994). Early occupations at Gran Pajatén, Peru. Andean Past 4, 281–318.

Clasby, R., & Meneses, J. (2012). Nuevas investigaciones en Huayurco: Resultados ini­
ciales de las excavaciones de un sitio de la ceja de selva de los Andes peruanos. Arque­
ología y Sociedad, 25, 303–326.

Page 28 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Clay, J. W., Alcorn, J. B., & Butler, J. R. (2000). Indigenous peoples, forestry management
and biodiversity conservation. Washington, DC: WWF and the World Bank.

Coetzee, B. W. T., Gaston, K. J., & and Chown, S. L. (2014). Local scale comparisons of
biodiversity as a test for global protected area ecological performance: A meta-
analysis. PlosOne, 9(8), e105824.

Commission on Development and Environment for Amazonia. (1992). Amazonia without


Myths. New York, NY: IADB/UNDP/Amazon Cooperation Treaty.

Commission on Development and Environment for Amazonia. (2001). Amazonia without


myths. Washington, DC: Inter-American Development Bank.

Costa, M. L., Kern, D. C., Pinto, A. H. Eleotério, & Souza, J. R. T. (2004). The ceramic arti­
facts in archaeological black earth (terra preta) from lower Amazon region, Brazil: Miner­
alogy. Acta Amazonica, 34(2), 165–178.

Costanza, R., D’Arge, R., de Groot, R., Farber, S., Grasso, M., Hannon, B., . . . van den
Belt, M. (1997). The value of the world’s ecosystem services and natural capital. Nature,
387, 253–260.

Cruz-Basso, L., & Kimura, H. (2006). The black earth eco-system recreating what
used to exist: The black earth of the Amazon Indians. SSRN, February 18.

Crystal, N., McMichael, H., Matthews-Bird, F., Farfan-Rios, W., & Feeley, K. J. (2017). An­
cient human disturbances may be skewing our understanding of Amazonian
forests. PNAS, 114(3), 522–527.

Dammert, J. L. (2013). Expansión de palma aceitera en la Amazonía: En las puertas del


escándalo. Lima, La Revista Agraria, 153, 4–5.

Damschen, E. I., Haddad, N. M., Orrock, J. L., Tewksbury, J. J., & Levey, D. J. (2006). Corri­
dors increase plant species richness at large scales. Science, 313(5791), 1284–1286.

Daniels, A. E. (2002). Indigenous peoples and neotropical forest conservation: Im­


pacts of protected area systems on traditional cultures. Macalester Environmental
Review, September 23.

DeFries, R., Hansen, A., Newton, A. C., & Hansen, M. C. (2005). Increasing isolation of
protected areas in tropical forests over the past twenty years. Ecological Applications, 15,
19–26.

Denevan, W. M. (1970). Aboriginal drained field cultivation in the Americas. Science, 169,
647–654.

Denevan, W. M. (1976). The aboriginal population of Amazonia. In W. Denevan (Ed.), The


native population of the Americas in 1492 (pp. 205–234). Madison: University of Wiscon­
sin Press.

Page 29 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Denevan, W. M. (1992). The pristine myth: The landscape of the Americas in 1492. Annals
of the Association of American Geographers, 82(3), 369–385.

Denevan, W. M. (2003). The Native Population of Amazonia in 1492 Reconsidered. Revista


de Indias, 63(227), 175–188.

Dent, D. H., & Wright, S. J. (2009). The future of tropical species in secondary forests: A
quantitative review. Biological Conservation, 142, 2833–2843.

Diegues, A. C. (2005). O mito moderno da natureza intocada. São Paulo, Brazil: Hucitec.

Donovan, W. (2012). The devastating costs of the Amazon Gold Rush. Smithsonian
Magazine, February.

Dórea, J. G., & Marques, R. C. (2016). Mercury levels and human health in the Ama­
zon basin. Annals of Human Biology, 43(4), 349–359.

Dorst, J. (1970). Before nature dies. London, U.K.: Collins.

Dourojeanni, M. J. (1968). Estado actual de la conservación de la flora y de la fauna en el


Perú. Ciencia Interamericana, 9(1–6), 1–12.

Dourojeanni, M. J. (1976). Machu Picchu and Peru’s national system of conservation units.
Washington DC Parks, 1(2), 8–11.

Dourojeanni, M. J. (1982). Recursos naturales y desarrollo en América Latina y el Caribe.


Lima, Peru: Universidad de Lima Press.

Dourojeanni, M. J. (1985a). An example of the complexity of the development in the humid


tropics: The Northwest Development Program in Brazil. University of Toronto Report to
IUCN.

Dourojeanni, M. J. (1985b). Overexploited and underutilized animals in the Amazon re­


gion. In G. T. Prance & T. E. Lovejoy (Eds.), Amazonia (pp. 419–433). New York, NY: Perga­
mon Press.

Dourojeanni, M. J. (1987). Aprovechamiento del barbecho forestal en áreas de agricultura


migratoria en la Amazonia Peruana. Lima, Revista Forestal del Perú, 14(2), 15–61.

Dourojeanni, M. J. (1990). Amazonia ¿Qué hacer? Iquitos, Peru: Centro de Estudios


Teológicos de la Amazonía (CETA).

Dourojeanni, M. J. (1992). Environmental impact of coca cultivation and cocaine produc­


tion in the Amazon region of Peru. Bulletin on Narcotics, United Nations, 44(2), 37–53.

Dourojeanni, M. J. (1999). The future of Latin American’s natural forests. In K. Keipi (Ed.),
Forest resource policy in Latin America (pp. 79–92). Washington, DC: Inter-American De­
velopment Bank and John Hopkins University Press.

Page 30 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Dourojeanni, M. J. (2002). Political will for establishing and managing parks. In J. Ter­
borgh, C. van Schaik, L. Davenport, & M. Rao (Eds.), Making parks work (pp. 320–334).
Washington, DC: Island Press.

Dourojeanni, M. J. (2006a). Estudio de caso sobre la carretera Interoceánica en la Amazo­


nia sur del Perú. Lima, Peru: Bank Information Center/Conservation International/Dere­
cho, Ambiente y Recursos Naturales.

Dourojeanni, M. J. (2006b). ¿Organizaciones no gubernamentales internacionales o


“transnacionales”? Lima, Ecología Aplicada, 5(1–2), 157–166.

Dourojeanni, M. J. (2008). Manejo comunitario de bosques tropicales y cooperación inter­


nacional: Lecciones no aprendidas. Bois et Forêts des Tropiques, 295(1), 47–57.

Dourojeanni, M. J. (2009a). Agroforestry systems and the environment. Amazon Agro­


forestry, 1(2), 3–4.

Dourojeanni, M. J. (2009b). Crónica forestal del Perú. San Marcos, Guatemala: Universi­
dad Nacional Agraria La Molina.

Dourojeanni, M. J. (2011a). Amazonia probable y deseable: Ensayo sobre el presente y el


futuro de la Amazonia. Lima, Peru: Universidad Inca Garcilazo de la Vega, Textos Univer­
sitarios.

Dourojeanni, M. J. (2011b). Evolución del financiamiento internacional en la Amazonia:


De Rondonia (Brasil) a Madre de Dios (Perú). Revista Latinoamericana de Derecho y
Políticas Ambientales, Lima, 1(1), 199–214.

Dourojeanni, M. J. (2013b). Loreto sostenible al 2021. Lima, Peru: Derecho, Ambiente y


Recursos Naturales.

Dourojeanni, M. J. (2015). Human impact on protected areas of the Peruvian Amazon. In


G. Wuerther, E. Crist, & T. Butler (Eds.), Protecting the Wild, Parks and Wilderness, the
Foundation for Conservation (pp. 215–225). Washington, DC: Island Press.

Dourojeanni, M. J. (2018). Amazonia ¿Qué futuro? Lima, Peru: Universidad Nacional de


Educación Guzman y Valle/ Grijley.

Dourojeanni, M. J., Barandiaran, A., & Dourojeanni, D. (2010). Amazonía Peruana en


2021: Explotación de recursos naturales e infraestructura (2nd ed.). Lima, Peru: ProNatu­
raleza/Derecho, Ambiente y Recursos Naturales/Iniciativa de Conservación de la Ama­
zonía/Sociedad Peruana de Derecho Ambiental.

Dourojeanni, M. J., & Padua, M. T. J. (2007). Biodiversidade: A hora decisiva (2nd ed.) Cu­
ritiba, Brazil: Editora UFPR.

Dourojeanni, M. J., & Padua, M. T. J. (2013). Arcas à deriva: Unidades de conservação do


Brasil. Rio de Janeiro, Brazil: Technical Books.

Page 31 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Dourojeanni, M. J., & Ponce, C. (1978). Los parques nacionales del Perú. Colección “La
Naturaleza en Iberoamérica.” Madrid, Spain: España Instituto de la Caza Fotógrafica (IN­
CAFO).

Dourojeanni, M. J., & Quiroga, R. E. (2006). Gestión de áreas protegidas para la conser­
vación de la biodiversidad: Evidencias de Brasil, Honduras y Perú. Washington, DC: De­
partamento de Desarrollo Sostenible, BID.

Dowie, M. (2005). Conservation refugees: When protecting nature means kicking people
out. Orion, 24, 16–27.

Duarte-Cardoso G., & do S. C. de Sousa Cardoso, A. (2013). Os reflexos da globalização


na governança da Amazônia: um estudo sobre os programas e estratégias para a defesa
de sua soberania. Anais dos Encontros nacionais da ANPUR, 15.

Dubois, J. C. L., Viana, V. M., & Anderson, A. (1996). Manual agroflorestal para a Amazô­
nia, Vol. 1. Rio de Janeiro, Brazil: Instituto Rede Brasileira Agroflorestal (REBRAF).

Dudley, N., & Stolton, S. (Eds.). (2007). Defining protected areas: An international confer­
ence in Almeria, Spain. Gland, Switzerland: IUCN, WCPA.

Edwards, D. P., Tobias, J. A., Sheil, D., Meijaard, E., & Laurance, W. F. (2014). Maintaining
ecosystem function and services in logged tropical forests. Trends in Ecology & Evolution,
20, 1–10.

Emrich, T. A., Pokorny, B., & Sepp, C. (2000). The significance of secondary forest man­
agement for development policy. Deutsche Gesellschaft für Technische Zusammenarbeit
(GTZ) GmbH Tropenökologisches Begleitprogramm, Eschborn.

Enrique, C., & Cueto, V. (2011). Propuestas para construir gobernanza en la Ama­
zonía a través del transporte sostenible revista. Latinoamericana de Derecho y Políti­
cas Ambientales, 1, 325–336.

Erickson, C. (2003). Historical ecology and future explorations. In J. Lehmann, D. Kern, B.


Glaser, & W. Woods (Eds.), Amazonian dark earths: Origin, properties, management (pp.
455–502). Dordrecht: Kluwer Academic Publishers.

Esquivel‐Muelbert, A., Baker, T. R., Dexter, K. G., Lewis, S. L., Brienen, R. J. W., Feld­
pausch, T. R., . . . Phillips, O. L. (2018). Compositional response of Amazon forests to
climate change. Global Change Biology, 25(1), 39–56.

Fairhead, J., Leach, M., & Scoones, I. (2011). Green grabbing: A new appropriation of na­
ture? Journal of Peasant Studies, 39(2), 237–261.

Fajardo, J., Lessmann, J., Bonaccorso, E., Devenish, Ch., & Munoz, J. (2014). Combined
use of systematic conservation planning, species distribution modelling, and connectivity

Page 32 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

analysis reveals severe conservation gaps in a megadiverse country (Peru). PlosOne, De­
cember 5.

Farber, S. C., Costanza, R., & Wilson, M. A. (2002). Economic and ecological concepts for
valuing ecosystem services. Ecological Economics, 41, 375–392.

Fearnside, P. M. (1985). Brazil’s Amazon forest and the global carbon problem. Intercien­
cia, 10(40), 179–186.

Fearnside, P. M. (1986). Settlement in Rondonia and the token role of science and tech­
nology in Brazil’s Amazonian development planning. Interciencia, 11, 229–236.

Fearnside, P. M. (1987). Rethinking continuous cultivation in Amazonia. Bioscience, 37(3),


209–213.

Fearnside, P. M. (1989). A ocupação humana de Rondônia: Impatos, límites e planejamen­


to. Programa Polonoroeste INPA, Manaus SCT/PR/CNPq Relatório de Pesquisa No. 5.

Fearnside, P. M. (1995). Agroforestry in Brazil’s Amazonian development policy: The role


and limits of a potential use for degraded lands. In M. Clusener-Godt & I. Sachs (Eds.),
Brazilian perspectives on sustainable development of the Amazon region (pp. 125–148).
Paris, France: UNESCO.

Fearnside, P. M. (2001). Environmental impacts of Brazil’s Tucuruí dam: Unlearned


lessons for hydroelectric development in Amazonia. Environmental Management, 27(3),
377–396.

Fearnside, P. M. (2005). Deforestation in Brazilian Amazonia: History, rates, and conse­


quence. Conservation Biology, 19(3), 680–688.

Fearnside, P. M. (2006). Balbina dam, Amazonas. Department of Ecology, National Insti­


tute for Research in the Amazon (INPA): Manaus.

Fearnside, P. M. (2007). Brazil’s Cuiabá-Santarém (BR-163) highway: The environmental


cost of paving a soybean corridor through the Amazon. Environmental Management, 39,
601–614.

Fearnside, P. M. (2008). Amazon forest maintenance as a source of environmental ser­


vices. Anais da Academia Brasileira de Ciências, 80(1), 101–114.

Fearnside, P. M. (2012). The Jirau dam’s CDM proposal: Comments on the project
design document. International Rivers, June 1.

Fearnside, P. M. (2014). Conservation research in Brazilian Amazonia and its contribution


to biodiversity maintenance and sustainable use of tropical forests. Paper presented at
First Conference on Biodiversity in the Congo Basin, Consortium Congo 2010, Université
de Kisangani, Kisangani, Democratic Republic of Congo, June 6–10.

Page 33 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Fearnside, P. M. (2015). Emissions from tropical hydropower and the IPCC. Environmen­
tal Science & Policy, 50, 225–239.

Fearnside P. M., & Pueyo, S. (2012). Greenhouse-gas emissions from tropical dams. Na­
ture Climate Change, 2, 382–384.

Fearnside, P. M., & Vilela, P. (2018). Chainsaw massacre: Protected areas in danger
in Brazil’s state of Rondônia. Mongabay, October 31.

Fernández, L.(2010). Desarrollo territorial en Madre de Dios: Los impactos socioambien­


tales de la Carretera Interoceánica Sur. Lima, Peru: Pronaturaleza.

Finner, M., Jenkins, C. N., Pimm, S. L., Keane, B., Ross, Y C. (2008). Oil and gas projects
in the Western Amazon: Threats to wilderness, biodiversity and indigenous people.
PlosOne, 3(8), e2932.

Finer, M., & Jenkins, C. N. (2012). Proliferation of hydroelectric dams in the Andean
Amazon and implications for Andes-Amazon connectivity. PlosOne, April 18.

Finer, M., Jenkins, C. N., Blue Sky, M. A., & Pine, J. (2014). Logging concessions enable
illegal logging crisis in the Peruvian Amazon. Scientific Reports, 4, 4719.

Finer, M., & Orta-Martínez, M. (2010). A second hydrocarbon boom threatens the Pe­
ruvian Amazon: Trends, projections, and policy implication. Environmental Re­
search Letters, 5(1), 1–10.

Finer, M., & Novoa, S. (2017). New gold mining deforestation zone: Upper Mali­
nowski (Madre de Dios, Peru). Monitoring of the Andean Amazon Project, MAAP: 72.

Finer, M., & Villa, L. (2018). New deforestation threats in the Peruvian Amazon
(Part 2: Agriculture expansion). Monitoring of the Andean Amazon Project, MAAP: 92.

Fitkau, E. J., & Klinge, H. (1973). On biomass and trophic structure of the Central Ama­
zonian rain forest. Biotropica, 5(1), 2–15.

Foley, J. A., Asner, G. P., Costa, M. H., Coe, M. T., DeFries, R., Gibbs, H. K., . . . Snyder, P.
(2007). Amazonia revealed: Forest degradation and loss of ecosystem goods and services
in the Amazon basin. Frontiers in Ecology and the Environment, 5, 25–32.

Food and Agriculture Organization–European Commission. (2012). Global forest land-use


change 1990–2005. FAO Forestry Paper No. 169, Rome.

Forsberg, B. R., Melack, J. M., Dunne, T., Barthem, R. B., Goulding, M., Paiva, R. C. D., . . .
Weisser, S. (2017). The potential impact of new Andean dams on Amazon fluvial
ecosystems. PlosOne, 12(8), e0182254.

Frasier, C. L., Albert, V. A., & Struwe, L. (2008). Amazonian lowland, white sand areas
as ancestral regions for South American biodiversity: Biogeographic and phylo­

Page 34 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

genetic patterns in Potalia (Angiospermae: Gentianaceae). Elsevier Science Direct


Organisms Diversity & Evolution, 8(1), 44–57.

Freire, W. I., Fonseca, R. L., do Prado Pereira, P. G., de Almeida Prado, A. C., Pimenta
Ribeiro, A., les Viana, E. M., . . . Panciera, E. F. (2007). Implicações da Iniciativa de Inte­
gração da Infra-estrutura Regional Sul-americana e projetos correlacionados na política
de conservação no Brasil. Conservation International Política Ambiental No. 3.

Galvin, M., & Haller, T. (Eds.). (2008). People, protected areas and global change partici­
patory conservation in Latin America, Africa, Asia and Europe. Bern, Switzerland: Swiss
National Centre of Competence in Research North-South University of Bern.

Garcia, J. G. (2014). Open government: Transparency, participation and collabora­


tion in public administration. Innovar, 24(54), 75–88.

Garnett, S. T., Burgess, N. D., Fa, J. E., Fernández-Llamazares, Á., Molnár, Z., Robinson, C.
J., . . . Leiper, I. (2018). A spatial overview of the global importance of indigenous lands
for conservation. Nature Sustainability Analysis, 1, 369–374.

Ghimire, K. B., & Pimbert, M. P. (Eds.). (1997). Social change and conservation. London,
U.K.: Earthscan.

Gibbs, H. K., Munger, J., L’Roe, J., Barreto, P., Pereira, R., Christie, M., . . . Walker, N. F.
(2015). Did ranchers and slaughterhouses respond to zero-deforestation agree­
ments in the Brazilian Amazon?Conservation Letters, 9(1), 32–42.

Gibson, L., Lee, T. M., Koh, L. P., Brook, B. W., Gardner, T. A., Barlow, J., . . . Sodhi, N. S.
(2011). Primary forests are irreplaceable for sustaining tropical biodiversity. Nature, 478,
378–381.

Gilbert, N. (2012). Palm-oil boom raises conservation concerns. Nature, News in Focus,
487, 14–15.

Glave, M., Hopkins, A., Malky, A., & Fleck, L. (2012). Análisis económico de la carretera
Pucallpa: Cruzeiro do Sul. GRADE, Avances de Investigación No. 4, Lima, Peru.

Goetz, S. J., Hansen, M., Houghton, R. A., Walker, W., Laporte, N., & Busch, J. (2015).
Measurement and monitoring needs, capabilities and potential for addressing re­
duced emissions from deforestation and forest degradation under REDD+. Envi­
ronmental Research Letters, 10(12).

Gómez-Pompa, A., & Kahm, A. (1992). Taming the wilderness myth. BioSciences, 42(4),
271–279.

Gómez-Pompa, A., Vasquez-Yañez, C., & Guevara, S. (1972). The tropical rain forest: A
non-renewable resource. Science, 177, 762–765.

Page 35 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Goodland, R. (Ed.). (2005). Oil and gas pipelines: Social and environmental impact assess­
ment. State of the Art. Fargo, IAIA 2005 Conference International Association of Impact
Assessment.

Goodland, R. J. A., & Irving, H. S. (1975). Amazon jungle: Green hell to red desert? New
York, NY: Elsevier.

Goulding, M. (1981). The fishes and the forest: Explorations in Amazonian Natural Histo­
ry. Berkeley: University of California Press.

Goulding, M., Smith, N. J. H., & Mahar, D. J. (1995). Floods of fortune: Ecology and econo­
my along the Amazon. New York, NY: Columbia University Press.

Grupo Regional Sobre Financiamiento e Infraestructura. (2014). Panorama del finan­


ciamiento para infraestructura en América Latina ¿Cuál es el contexto regional en el que
se inserta el nuevo Banco de los BRICS? Lima, Peru: DAR.

Gutierrez-Velez, V. H., DeFries, R., Pinedo-Vasquez, M., Iriarte, M., Padoch, Ch., Baeth­
gen, W., . . . Lim, Y. (2011). High-yield oil palm expansion spares land at the ex­
pense of forests in the Peruvian Amazon. Environmental Research Letters, 6, 044029.

Haddad, N. M., Brudvig, L. A., Clobert, J., Davies, K. F., Gonzalez, A., Holt, R. D., . . .
Townshend, J. R. (2015). Habitat fragmentation and its lasting impact on earth’s
ecosystems. Science Advances, 1(2), e1500052.

Haddad, N. M., Brudvig, L. A., Damschen, E. I., Evans, D. M., Johnson, B. L., Levey, D.
J., . . . Weldon, A. J. (2014). Potential negative ecological effects of corridors. Conservation
Biology, 28, 1178–1187.

Hamilton, M. J., Walker, R. S., & Kesler, D. C. (2014). Crash and rebound of indigenous
populations in lowland South America. Scientific Reports, 4, 4541.

Hansen, M. C., Potapov, P. V., Moore, R., Hancher, M., Turubanova, S. A., Tyukavina,
A., . . . Townshend, J. R. G. (2013). High-resolution global maps of 21st-century forest cov­
er change. Science, 342, 850–853.

Harris, N. L., Brown, S., Hagen, S. C., Saatchi, S. S., Petrova, S., Salas, W., . . . Lotsch, A.
(2012). Baseline map of carbon emissions from deforestation in tropical regions. Science,
336(22), 1573–1576.

Hartshorn, G. (1980). Neotropical forest dynamic. Biotropica, 12, 23–30.

Hecht, S. B. (1985). Environment, development and politics: Capital accumulation and the
livestock sector in the eastern Amazon. World Development, 13(6), 663–684.

Hecht, S. B. (1993). The logic of livestock and deforestation in Amazonia. Bioscience, 43,
687.

Page 36 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Hecht, S. B., Norgaard, R., & Possio, G. (1988). The economics of cattle ranching in east­
ern Amazonia. Interciencia, 13, 233–240.

Holdridge, L. R. (1947). Ecología basada en zonas de vida. San José, Costa Rica: IICA.

Houghton, R. A. (2003). Revised estimates of the annual net flux of carbon to the atmos­
phere from changes in land use and land management. Tellus, 55, 378–390.

Houghton, R. A., Skole, D. L., Nobre, C. A., Hackler, J. L., Lawrence, K. T., & Chomen-
Towski, W. H. (2000). Annual fluxes of carbon from deforestation and regrowth in the
Brazilian Amazon. Nature, 403(6767), 301–304.

International Tropical Timber Organization. (2006). Status of tropical forest management


2005. Yokohama, Japan: International Tropical Timber Organization.

International Tropical Timber Organization–United Nations University. (2011). Trans­


boundary conservation and peace-building: Lessons from forest projects. Yokohama,
Japan: International Tropical Timber Organization.

International Union for Conservation of Nature. (1971). United Nations list of national
parks and equivalent reserves. IUCN New Series No. 15.

International Union for Conservation of Nature. (1990). 1990 United Nations list of na­
tional parks and protected areas. Gland, Switzerland: World Conservation Union.

Izko, X. (2012). La frontera invisible: Actividades extractivas, infraestructura y ambiente


en la Amazonia Ecuatoriana 2010–2030. Quito, Ecuador: ICAA.

Jakovac, C. C., Pena-Claros, M., Kuyper, Th. W., & Bongers, F. (2015). Loss of secondary-
forest resilience by land-use intensification in the Amazon. Journal of Ecology, 103, 67–77.

Kalamandeen, M., Gloor, E., Mitchard, E., Quincey, D., Ziv, G., Spracklen, D., . . . Gal­
braith, D. (2018). Pervasive rise of small-scale deforestation in Amazonia. Scientific
Reports, 8, 1600.

Kammesheidt, L., Köhler, P., & Huth, A. (2002). Simulating logging scenarios in sec­
ondary forest embedded in a fragmented neotropical landscape. Forest Ecology
and Management, 170(1), 89–105.

Kareiva, P., & Marvier, M. (2003). Conserving biodiversity cold spots. American Scientist,
91, 344–351.

Kareiva, P., Marvier, M., & Lalasz, R. (2012). Conservation in the Anthropocene: Be­
yond solitude and fragility. Breakthrough Journal.

Kellert, S. R., Mehta, J. N., Ebbin, S. A., & Lichenfeld, L. L. (2000). Community natural re­
source management: Promise, rhetoric and reality. Society & Natural Resources, 13, 705–
715.

Page 37 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Kim, D. H., Sexton, J. O., & Townshend, J. R. (2015). Accelerated deforestation in the hu­
mid tropics from the 1990s to the 2000s. Geophysical Research Letters, 42, 3495–3501.

Klinge, H., Rodriguez, W. A., Bruning, E., & Fitkau, E. J. (1975). Biomass and structure in
a central Amazonian rain forest. In E. F. Golley & E. Medina (Eds.), Tropical ecological
system (pp. 115–122). Berlin Ecological Studies 11. Berlin, Germany: Springer Verlag.

Koren, Y., Kaufman, J., Remer, L. A., & Martins, J. V. (2004). Measurement of the effect of
Amazon smoke on inhibition of cloud formation. Science, 303(5662), 1342–1345.

Lahteenoja, O., Rojas Reategui, Y., Rasanen, M., del Castillo Torres, D., Oinonen, M., &
Page, S. (2012). The large Amazonian peatland carbon sink in the subsiding Pas­
taza-Maranon foreland basin, Peru. Global Change Biology, 18, 164–178.

Lathrap, D. W. (1973). The antiquity and importance of long distance trade relationships
in the moist forest of pre-Columbian South America. World Archeology, 5(2), 170–186.

Laurance, W. F., Camargo, J. L. C., Luizão, R. C. C., Laurance, S. G., Pimm, S. L., Bruna, E.
M., . . . Lovejoy, T. E. (2011). The fate of Amazonian forest fragments: A 32-year investiga­
tion. Biological Conservation, 144, 56–67.

Laurance, W. F., Lovejoy, T. E., Vasconcelos, H. L., Bruna, E. M., Didham, R. K., Stouffer,
P., . . . Sampiao, E. (2002). Ecosystem decay of Amazonian forest fragments: A 22-year in­
vestigation. Conservation Biology, 16(3), 605–618.

Laurance, W. F., Oliveira, A. A., Laurance, S. G., Condit, R., Nascimento, H. E. M.,
Sanchez-Thorin, A. C., . . . Dick, C. W. (2004). Pervasive alteration of tree communities in
undisturbed Amazonian forests. Nature, 428, 171–175.

Laurance, W. F., Goosem, M., Laurance, S. G. W. (2009). Impacts of roads and linear clear­
ings on tropical forests. Trends in Ecology & Evolution, 24, 659–669.

Laurance, W. F., Sayer, J., & Cassman, K. G. (2014). Agricultural expansion and its impacts
on tropical nature. Trends in Ecology & Evolution, 29(2), 107–118.

Laurance, W. F., Useche, D. C., Rendeiro, J., Kalka, M., Bradshaw, C. J. A., Sloan, S. P., . . .
Zamzani, F. (2012). Averting biodiversity collapse in tropical forest protected areas. Na­
ture, 489, 290–294.

Laurance, W. F., Vasconcelos, H. L., & Lovejoy, T. E. (2001). Forest loss and fragmentation
in the Amazon: Implications for wildlife conservation. Oryx, 34(1), 39–45.

Leal-Farias, D. B. (2018). Aid and technical cooperation as a foreign policy tool for emerg­
ing donors: The case of Brazil. London, U.K.: Routledge Explorations in Development
Studies.

Lentini, M., Pereira, D., Celentano, D., & Pereira, R. (2005). Fatos florestais da Amazônia.
Belém, Brazil: Imazon.

Page 38 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

León, E. (2013). 14,000 años de alimentación en el Perú. Lima, Peru: Universidad San
Martin de Porres.

León, F. (2007). El aporte de las áreas naturales protegidas a la economía nacional. Lima,
Peru: INRENA.

Levis, C. C., Costa, F. R. C., Bongers, F., Peña-Claros, M., Clement, C. R., Junqueira, A.
B., . . . Ter Steege, H. (2017). Persistent effects of pre-Columbian plant domestication on
Amazonian forest composition. Science, 355, 925–930.

Little, P. E. (2013). Megaproyectos en la Amazonia: Un análisis geopolítico y socioambien­


tal con propuestas de mejor gobierno para la Amazonía. Lima, Peru: Derecho Ambiente y
Recursos Naturales (DAR).

Lovejoy, T. E. (2015). Ask an Amazon expert: Why we can’t afford to lose the rain
forest. National Geographic, May 19.

Lovejoy, T. E., & Nobre, C. (2018). Amazon tipping point. Science Advances, 4(2).

Lugo, A. E. (2013). Novel tropical forests: Nature’s response to global change. Special is­
sue, Tropical Conservation Science, 6(3), 325–337.

Lumbreras, L. G. (1981). Arqueología de la América Andina. Lima, Peru: Milla Batres.

Lumbreras, L. G. (2013). Los orígenes de la civilización en el Perú. Cusco, Peru: Ministe­


rio de Cultura, Subdirección Desconcentrada de Cultura del Cusco.

Lund, H. G. (2009). What is a degraded forest? White Paper prepared for FAO, UN Food,
and Agriculture Organization, Rome.

MacDicken, K. G., & Tubiello, F. N. (2015). Seeing the trees but not the forest. Mongabay,
March 20.

Machlis, G. E. (1992). The contribution of sociology to biodiversity research and manage­


ment. Biological Conservation, 62, 161–170.

Machlis, G. E., & Tichnell, D. L. (1985). The state of the world’s parks: An international
assessment of resources management. Boulder, CO: Westview Press.

Makarieva, A. M., & Gorshkov, V. G. (2007). Biotic pump of atmospheric moisture as dri­
ver of the hydrological cycle on land. Hydrology and Earth System Sciences, 11, 1013–
1033.

Makarieva, A. M., Gorshkov, V. G., Sheil, D., Nobre, A. D., Bunyard, P., & Li, B.-L. (2014).
Why does air passage over forest yield more rain? Examining the coupling between rain­
fall, pressure, and atmospheric moisture content. Journal of Hydrometeorology, 15, 411–
426.

Page 39 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Malky, A., Reid, J., Ledezma, J. C., & Fleck, L. (2011). El filtro de carreteras: Un análisis
estratégico de proyectos viales en la Amazonía. Conservation Strategy Fund, Série Técni­
ca No. 21.

Mann, C. (2011). 1493: Uncovering the New World Columbus created. New York, NY: A.
Knopf.

Marengo, J. A., Tomasella, J., Soares, W. R., Alves, L. M., & Nobre, C. A. (2011). Extreme
climatic events in the Amazon basin. Theoretical and Applied Climatology, 107, 73–85.

Margulis, S. (2004). Causes of deforestation of the Brazilian Amazon. World Bank Work­
ing Paper No. 22, Washington, DC.

Martins de Souza, J. (1996). O tempo da fronteira: Retorno à controvérsia sobre o tempo


histórico da frente de expansão e da frente pioneira. Tempo Social: Revista de Sociologia
da USP, 8(1), 25–70.

Mascia, M. B., Pailler, S., Krithivasan, R., Roschchanka, V., Burns, D., Mlotha, M. J., . . .
Peng, N. (2014). Protected area downgrading, downsizing, and degazettement (PADDD) in
Africa, Asia, and Latin America and the Caribbean, 1900–2010. Biological Conservation,
169, 355–361.

Mashuta, K., Ovando, A., & McKey, D. (2018). The unique functioning of a pre-Columbian
Amazonian floodplain fishery. Scientific Reports, 8, 5998.

McClain, M. E., Victoria, R. L., & Richey, J. E. (Eds.). (2001). Biochemistry of the Amazon
basin. Oxford, U.K.: Oxford University Press.

McFarland, W., Whitley, S., & Kissinger, G. (2015). Subsidies to key commodities driving
forest loss: Overseas Development Institute. Mongabay, March.

McMichael, C. N. H., Matthews-Bird, M. F., Farfan-Rios, W., & Feeley, K. J. (2017). Ancient
human disturbances may be skewing our understanding of Amazonian forests. PNAS,
114(3), 522–527.

Medina, M. H. (1980). Treaty for Amazonian cooperation: General analysis. In F. Barbira-


Scazzochio (Ed.), Land, people planning in contemporary Amazon (pp. 58–71). Cam­
bridge, U.K.: Centre of Latin American Studies.

Meggers, B. J. (1971). Amazonia: Man and culture in counterfeit paradise. Washington,


DC: Smithsonian.

Melián, C. J., Seehausen, O., Eguíluz, V. M., Fortuna, M. A., & Deiner, K. (2015). Diversifi­
cation and biodiversity dynamics of hot and cold spots. Ecography, 35(4), 393–401.

Miller, R. P., & Nair, P. K. R. (2006). Indigenous agroforestry systems in Amazonia:


From prehistory to today. Agroforestry Systems, 66(2), 151–164.

Page 40 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Millikan, B. H. (1984). The dialectics of devastation. Berkeley: University of California


Press.

Ministerio del Ambiente Peru. (2010). Manual de operaciones del Programa Nacional de
Conservación de Bosques. Lima, Peru: Ministerio del Ambiente/Programa Nacional de
Conservación de Bosques.

Ministério do Meio Ambiente. (2001). Avaliação e identificação de ações prioritárias para


a conservação, utilização sustentável e repartição dos benefícios da biodiversidade na
Amazônia Brasileira. Ministério do Meio Ambiente, Brasília DF, Brazil.

Ministerio dos Transportes. (2013). Evolução da malha rodoviária Brasília.

Mittermeier, R. A., & Bowles, I. A. (1993). The Global Environment Facility and biodiversi­
ty conservation: Lessons to date and suggestions for future action. Biodiversity and Con­
servation, 2, 637–655.

Mittermeier, R. A., Myers, N., Thomsen, J. B., da Fonseca, G. A. B., & Olivieri, S. (1998).
Biodiversity hotspots and major tropical wilderness areas: Approaches to setting conser­
vation priorities. Conservation Biology, 12(3), 516–520.

Mittermeier, R. A., Robles, P., & Mittermeier, C. C. (1997). Megadiversity. Mexico: CE­
MEX/Conservation International.

Mora, C., & Sale, P. F. (2011). Ongoing global biodiversity loss and the need to move be­
yond protected areas: A review of the technical and practical shortcomings of protected
areas on land and sea. Marine Ecology Progress Series, 434, 251–266.

Morim Novaes, R. L., & de França Souza, R. (2013). Legalizing environmental exploita­
tion in Brazil: The retreat of public policies for biodiversity protection. Tropical Conserva­
tion Science, 6(4), 477–483.

Murphree, M. W. (2002). Protected areas and the commons. Common Property Resource
Digest, 60, 1–3.

Myers, N. (1984). The primary source: Tropical forests and our future. New York, NY:
Norton.

Myers, N. (1988). Threatened biotas: “Hot spots” in tropical forests. The


Environmentalist, 8, 1–2.

Nasi, R., Taber, A., & van Vliet, N. (2011). Empty forests, empty stomachs? Bushmeat and
livelihoods in the Congo and Amazon basin. International Forestry Review, 13(3), 355–
387.

National Research Council. (1982). Ecological aspects of development in the humid trop­
ics. Washington, DC: National Academy Press.

Page 41 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Nepstad, D. (1997). Empobrecimento biológico da floresta amazônica por seringueiros,


madeireiros e fazendeiros. In T. Ximemes (Ed.), Perspectivas do desenvolvimento susten­
tavel (pp. 311–359). Belém, Brazil: Universidade Federal do Pará.

Nepstad, D., Carvalho, G., Barros, A. C., Alencar, A., Capobianco, J. P., Bishop, J.,. . . Prins,
E. (2001). Road paving, fire regime feedbacks, and the future of Amazon forests. Forest
Ecology and Management, 154, 395–407.

Nepstad, D., Lefebvre, P., Lopes da Silva, U., Tomasella, J., Schlesinger, P., Solorzano,
L., . . . Guerreira Benito, J. (2004). Amazon drought and its implications for forest flamma­
bility and tree growth: A basin-wide analysis. Global Change Biology, 10, 704–717.

Nepstad, D., Lefebvre, P., Lopes da Silva, U., Tomasella, J., Schlesinger, P., Solorzano,
L., . . . Fisher, J. (2009). Drought sensitivity of the Amazon rainforest. Science, 80(323),
1344–1347.

Nepstad, D., Schwartzman, S., Bamberger, B., Santilli, M., Ray, D., Schlesinger, P., . . . Rol­
la, A. (2006). Inhibition of Amazon deforestation and fire by parks and indigenous lands.
Conservation Biology, 20(1), 65–73.

Nepstad, D., Soares-Filho, B. S., Merry, F., Lima, A., Moutinho, P., Carter, J., . . . Stella, O.
(2009). The end of deforestation in the Brazilian Amazon. Science, Policy Forum, 326,
1350–1351.

Nepstad, D. C., Verissimo, A., Alencar, A., Lima, E., Lefebvre, P., Schlesinger, P., . . .
Moutinho, P. (1999). Large-scale impoverishment of Amazonian forests by logging and
fire. Nature, 398(8), 505–508.

Newell, R., & Newell, N. (1992). Tropospheric rivers? A pilot study. Geophysical Research
Letters, 12, 2401–2404.

Newson, L. A. (1996). The population of the Amazon basin in 1492: A view from the
Ecuadorian headwaters. Transactions of the Institute of British Geographers, 21(1), 5–26.

Nobre, A. D. (2014). O futuro climático da Amazônia. Relatório de Avaliação Científica


São José dos Campos, SP: ARA: CCST-INPE/INPA.

Olander, L. P., Gibbs, H. K., Steininger, M., Swenson, J. J., & Murray, B. C. (2008). Refer­
ence scenarios for deforestation and forest degradation in support of REDD: A
review of data and methods. Environmental Research Letters, 3(2).

Padua, M. T., & Coimbra-Filho, A. (1979). Os parques nacionais do Brasil. Madrid, Spain:
Incafo.

Pärssinen, M., Schaan, D., & Ranzi, A. (2009). Pre-Columbian geometric earthworks in
the upper Purús: A complex society in western Amazonia. Antiquity, 83(322), 1084–1095.

Page 42 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Peres, C. A. (2000). Evaluating the impact and sustainability of subsistence hunting at


multiple Amazonian sites. In J. G. Robinson & E. L. Bennett (Eds.), Hunting for sustain­
ability in tropical forests (pp. 31–57). New York, NY: Columbia University Press.

Peres, C. A. (2011). Conservation in sustainable-use tropical forest reserves. Conser­


vation Biology, 25(6), 1124–1129.

Peres, C. A., Baider, C., Zuidema, P. A., Wadt, L. H. O., Kainer, K., Gomes-Silva, D. A., . . .
Freckleton, R. P. (2003). Demographic threats to the sustainability of Brazil nuts exploita­
tion. Science, 302(5653), 2112–2114.

Peres, C. A., Gardner, T. A., Barlow, J., Zuanon, J., Michalski, F., Lees, A., . . . Feeley, K. J.
(2010). Biodiversity conservation in human-modified Amazonian forest landscapes. Bio­
logical Conservation, 143, 2314–2327.

Peres, C. A., & Terborgh, J. W. (1995). Amazonian nature-reserves: An analysis of the de­
fensibility status of existing conservation units and design criteria for the future. Conser­
vation Biology, 9, 34–46.

Perz, S. G., Overdevest, C., Caldas, M. M., Walker, R. T., & Arima, E. Y. (2007). Unofficial
road building in the Brazilian Amazon: Dilemmas and models for road governance. Envi­
ronmental Conservation, 34, 112–121.

Pierret, P. V., & Dourojeanni, M. J. (1966). La caza y la alimentación humana en las rib­
eras del Río Pachitea, Perú. Costa Rica, Turrialba, 16(3), 271–277.

Pierret, P. V., & Dourojeanni, M. J. (1967). Importancia de la caza para alimentación hu­
mana en el curso inferior del río Ucayali, Perú. Lima, Revista Forestal del Perú, 1(2), 10–
21.

Pinho, R. C., Miller, R. P., & Alfaia, S. (2012). Agroforestry and the improvement of
soil fertility: A view from Amazonia. Applied and Environmental Soil Science, 2012,
616383.

Plotkin, M. J. (1993). Tales of a shaman’s apprentice. London, U.K.: Penguin.

Pokorny, B., Sabogal, C., Prabhu, R., & Silva, J. N. M. (2002). Introducing criteria and
indicators for monitoring and auditing forest management in the Brazilian Ama­
zon. CIFOR and Embrapa Amazonia Oriental, Belem, Brazil.

Poole, P. (1989). Developing a partnership of indigenous peoples, conservationists, and


land use planners in Latin America. World Bank Policy, Planning and Research Working
Paper.

Porter-Bolland, L., Ellis, E. A., Guaraiguata, M. R., Ruiz-Mallen, I., Negrete, S., & Reyes,
V. (2012). Community managed forests and forest protected areas: An assessment of their
conservation effectiveness across the tropics. Forest Ecology Management, 268, 6–17.

Page 43 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Portugués, H., & Huerta, P. (2005). Mapa de deforestación de la Amazonía peruana 2000.
Lima, Peru: PROCLIM/CONAM.

Pöschl, U., Martin, S. T., Sinha, B., Chen, Q., Gunthe, S. S., Huffman, J. A., . . . Andreae, M.
O. (2010). Rainforest aerosols as biogenic nuclei of clouds and precipitation in the Ama­
zon. Science, 329, 1513–1516.

Post, M. (1990). The debt-for-nature swap: A long-term investment for the economic sta­
bility of less developed countries. International Lawyer, 24(4), 1071–1098.

Prance, Gh. T. (Ed.). (1982). Biological diversification in the tropics. New York, NY: Colum­
bia University Press.

ProInversión. (2016). Hidrovía Amazónica: Ríos Marañón y Amazonas, tramo


Saramiriza–Iquitos–Santa Rosa; Río Huallaga, tramo Yurimaguas–Confluencia
con el río Marañón; Río Ucayali, tramo Pucallpa–confluencia con el río Marañón.
Lima, Peru.

Putz, F. E., & Romero, C. (2012). Helping curb tropical forest degradation by linking
REDD+ with other conservation interventions: A view from the forest. Current Opinion in
Environmental Sustainability, 4, 1–8.

Putz, F. E., Sist, P., Fredericksen, T., & Dykstra, D. (2008). Reduced-impact logging: Chal­
lenges and opportunities. Forest Ecology and Management, 256(7), 1427–1433.

Putzel, L., Peters, C. M., & Romo, M. (2011). Post-logging regeneration and recruitment
of shihuahuaco (Dipteryx spp.) in Peruvian Amazonia: Implications for management. Ecol­
ogy and Management, 261, 1099–1105.

Quammen, D. (1996). National parks: Nature’s dead end. New York Times, July 28.

Ranzi, A., & Aguiar, R. (2004). Geoglyphs of Amazonia: Aerial perspective. Faculdade
de Energia, Universidade Federal de Santa Catarina, Florianópolis.

Red Amazónica de Información Socioambiental Georreferenciada. (2012). Amazonía bajo


presión.

Red Amazónica de Información Socioambiental Georreferenciada. (2015). Desmatamento


na Amazônia (1970–2013). Mapas e quadros estadísticos.

Red Amazónica de Información Socioambiental Georreferenciada. (2017). Amazonía 2017


áreas protegidas, territorios indígenas. Mapa y cuadros estadísticos.

Redford, K. H., & Padoch, C. (1992). Conservation of neotropical forests: Working from
traditional resource use. New York, NY: Columbia University Press.

Redford, K. H., & Stearman, A. M. (1993). Forest-dwelling native Amazonians and the
conservation of biodiversity. Conservation Biology, 7, 248–255.

Page 44 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Reid, W. V., & Miller, K. R. (1989). Keeping options alive: The scientific basis for conserv­
ing biodiversity. Washington, DC: World Resources Institute.

Reilly, W. (2006). Using international finance to further conservation: The first 15 years of
debt-for-nature swaps. In C. Jochnick & F. A. Preston (Eds.), Sovereign debt at the cross­
roads: Challenges and proposals for resolving the Third World debt crisis (pp. 197–214).
Oxford, U.K.: Oxford University Press.

Revelle, R. (1982). Carbon dioxide and world climate. Scientific American, 247(2), 33–41.

Riva, A. L. M. de, da Fonseca, L. F. L., & Hasenlever, L. (2008). Instrumentos econômi­


cos e financeiros para a conservação ambiental no Brasil. Instituto Socio-Ambiental
(ISA).

Rodrigues, E. P., & da Silva Pinheiro, E. (2011). O desflorestamento ao longo da


rodovia BR-174 (Manaus/AM–Boa Vista/RR). Sociedade & Natureza, 23(3), 513–528.

Rojas, M., & Castaño, C. (1990). Áreas protegidas de la cuenca del Amazonas. Bogotá,
Colombia: Tratado de Cooperación Amazónica.

Roosevelt, A. C. (2000). The lower Amazon: A dynamic human habitat. In D. L. Lentz


(Ed.), Imperfect balance: Landscape transformations in the Precolumbian Americas (pp.
455–491). New York, NY: Columbia University Press.

Roosevelt, A. C. (2013). Prehistory of Amazonia. In C. Renfrew & P. Bahn (Eds.), Cam­


bridge world prehistory (pp. 1175–1199). Cambridge, U.K.: Cambridge University Press.

Rosenberg, D. K., Noon, B. R., & Meslow, E. C. (1997). Biological corridor: Form, func­
tion, and efficacy. BioScience, 47, 677–687.

Rudel, T. K. (1983). Roads, speculators, and colonization in the Ecuadorian Amazon. Hu­
man Ecology, 11(4), 385–403.

Salati, E., Dall’Olio, A., & Matsui, J. R. G. (1979). Recycling of water in the Amazon basin:
An isotopic study. Water Resources Research, 15, 1250–1258.

Salati, E., & Vose, P. (1984). Amazon basin: A system in equilibrium. Science, 225(4568),
129–205.

Salo, J. (1987). Pleistocene forest refuges in the Amazon: Evaluation of the biostratigraph­
ical, lithostratigraphical and geomorphological data. Annales Zoologici Fennici, 24(3),
203–211.

Sanchez, P. A. (1976). Properties and management of soils in the tropics. New York, NY:
Wiley.

Sanchez, P. A., Bandy, D. E., Villachica, J. H., & Nicholaides, J. J. (1982). Amazon basin
soils: Management for continuous crop production. Science, 216, 821–827.

Page 45 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Sanchez, P. A., & Buol, S. W. (1975). Soils of the tropics and the world food crisis. Science,
188, 821–827.

Sanchez-Cuervo, A. M., Aide, T. M., Clark, M. L., & Etter, A. (2012). Land cover change in
Colombia: Surprising forest recovery trends between 2001 and 2010. PlosOne, 7(8),
e43943.

Sanford, R. L., Jr., Saldarriaga, J., Clark, K. E., Uhl, C., & Herrera, R. (1985). Amazon
rain-forest fires. Science, 227(4682), 53–55.

San Sebastián, M., & Hurtig, A.-K. (2004). Oil exploitation in the Amazon basin of
Ecuador: A public health emergency. Revista Panamericana de Salud Pública, Wash­
ington, 15(3). 205–211.

Santilli, J. (2005). Socioambientalismo e novos direitos. São Paulo, Brazil: Editora Peirópo­
lis.

Sasaki, N., & Putz, F. E. (2009). Critical need for new definitions of “forest” and “forest
degradation” in global climate change agreements. Conservation Letters, 2(5), 226–232.

Sayre, R., Roca, E., Sedaghatkish, G., Young, B., Keel, S., Roca, R., & Sheppard, S. (1999).
Nature in focus: Rapid ecological assessment. New York, NY: Island Press.

Schmidt, R. (1987). Tropical rain forest management: A status report. Unasylva, 156(39),
2–17.

Schultes, R. E. (1979). The Amazonia as a source of new economic plants. Economic


Botany, 33, 259–266.

Schwartzman, B. S., Nepstad, D., & Moreira, A. (2000). Arguing tropical forest conserva­
tion: People versus parks. Conservation Biology, 14(5), 1370–1374.

Serra, J. (2010). Inambari: La urgencia de una discusión seria y nacional. Lima, Peru:
Pronaturaleza.

Shady, R. (2006). America’s first city? The case of Late Archaic Caral. In Andean archaeol­
ogy III (pp. 28–66). Boston, MA: Springer.

Shane, D. R. (1980). Hoof prints on the forest: An inquiry into the beef cattle industry in
the tropical forest areas of Latin America. Washington, DC: Department of State Office of
Environmental Affairs.

Sinovas, P., Price, B., King, E., Hinsley, A., & Pavitt, A. (2017). Wildlife trade in the Ama­
zon countries: An analysis of trade in CITES listed species. Technical report prepared for
the Amazon Regional Program (BMZ/DGIS/GIZ). Cambridge, U.K.: UN Environment–
World Conservation Monitoring Centre.

Small, G. (2017). The fight is on for South American Bioceanic railway. Executive In­
telligence Review, 44(44).
Page 46 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Smith, R. Ch. (1983). Las comunidades nativas y el mito del gran vacío amazónico. Lima,
Peru: AIDESEP.

Soares-Filho, B. S., Nepstad, D. C., Curran, L. M., Cerqueira, G. C., Garcia, R. A., Ramos,
C. A., . . . Schlesinger, P. (2006). Modelling conservation in the Amazon basin. Nature,
440(23), 520–523.

Soares-Filho, B. S., Rajão, R., Macedo, M., Carneiro, A., Costa, W., Coe, M., . . . Alencar, A.
(2014). Cracking Brazil’s forest code. Policy Forum, Science, 344, 363–364.

Somarriba, E. J., Beer, J., Jr., Alegre Show, J. C., Andrade, H. J., Cerda, R., DeClerck, F., . . .
Campos, J. J. (2012). Mainstreaming agroforestry in Latin America. In P. K. R. Nair &
D. Garrity (Eds.), Agroforestry: The future of global land use (pp. 429–453). Advances in
Agroforestry 9. Dordrecht, Netherlands: Springer.

Souza, C., Jr., & Roberts, D. (2005). Mapping forest degradation in the Amazon region
with Ikonos images. International Journal of Remote Sensing, 26, 425–429.

Steege, H., Pitman, N. C., Sabatier, D., Baraloto, C., Salomão, R. P., Guevara, J. E., . . . Sil­
man, M. R. (2013). Hyperdominance in the Amazonian tree flora. Science, 342(6156),
1243092.

Swenson, J. J., Carte, C. E., Domec, J.-Ch., & Delgado, C. (2011). Gold mining in the Peru­
vian Amazon: Global prices, deforestation, and mercury imports. PlosOne, 6(4), e18875.

Swenson, J. J., Young, B. E., Beck, S., Comer, P., Córdova, J. H., Dyson, J., . . . Zambrana-
Torrelio, C. M. (2012). Plant and animal endemism in the eastern Andean slope:
Challenges to conservation. BMC Ecology, 12(1).

Tello, J. C. (1960). Chavín: Cultura matriz de la civilización andina, Part 1, Vol. 2. Lima, Pe­
ru: Publicación antropológica del Archivo “Julio C. Tello” de la Universidad Nacional May­
or de San Marcos.

Terborgh, J. (1999). Requiem for nature. Washington, DC: Island Press.

Terborgh, J. (2000). The fate of tropical forests: A matter of stewardship. Conservation Bi­
ology, 14(5), 1358–1361.

Tickell, O. (2013). Amazon: 441 new species discovered in four years. Ecologist, Oc­
tober 23.

Tosi, J. (1960). Zonas de vida natural en el Perú. IICA Zona Andina Report No. 5, Lima, Pe­
ru.

Udvardy, M. D. F. (1975). A classification of the biogeographical provinces of the world.


Morges, Switzerland: IUCN.

Page 47 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Uriartea, M., Pinedo-Vasquez, M., DeFriesa, R. S., Fernandes, K., Gutierrez-Veleza, V.,
Baethgen, W. E., & Padoch, C. (2012). Depopulation of rural landscapes exacerbates fire
activity in the western Amazon. PNAS, 109(52), 21546–21550.

Urrunaga, J. M., Johnson, A., Orbegozo, I. D., & Mulligan, F. (2012). The laundering ma­
chine: How fraud and corruption in Peru’s concession system are destroying the future of
its forests. Washington, DC: Environmental Investigation Agency EIA-GLOBAL.ORG.

van Dijc, P. (2013). The impact of the IIRSA road infrastructure programme on Amazonia.
Exeter, U.K.: Routledge.

Vera-Diaz, M. del C., Kaufmann, R. K., & Nepstad, D. C. (2009). The environmental im­
pacts of soybean expansion and infrastructure development in Brazil’s Amazon
basin. Global Development and Environment Institute Working Paper No. 09–05.

Veríssimo, A., & Nussbaum, R. (2011). Um resumo do status das florestas em países sele­
cionados: Nota técnica. São Paulo, Brazil: Imazon and Proforest.

Veríssimo, A., Rolla, A., Vedoveto, M., & Futada, S. de M. (2011). Protected areas in the
Brazilian Amazon: Challenges and opportunities. Belém, Brazil: Imazon.

Visser, D. R., & Mendoza, G. A. (1994). Debt-for-nature swaps in Latin America. Journal of
Forestry, 92(6), 13–16.

Wade, L. (2014). Searching for the Amazon’s hidden civilizations. AAAS Science
News, January 7.

Walker, R. S., & Hamilton, M. J. (2014). Amazonian societies on the brink of extinction.
Journal of Human Biology, 26(4), 570–572.

Watling, J., Iriarte, J., Mayle, F. E., Schaan, D., Pessenda, L. C. R., Loader, N. J., . . . Ranzi,
A. (2017). Impact of pre-Columbian “geoglyph” builders on Amazonian forests.
Proceedings of the National Academy of Sciences, 114(8), 201614359.

Watson, J. E., Dudley, N., Segan, D. B., & Hockings, M. (2014). The performance and
potential of protected areas. Nature, 15(7525), 67–73.

Watters, R. F. (1971). La agricultura migratoria en América Latina. FAO Cuaderno Fore­


stal No. 17, Rome, Italy.

Webster, D., & Haviv, R. (2012). The devastating costs of the Amazon Gold Rush.
Smithsonian Magazine, February.

Wetterberg, G. B. (2004). The concept of protected areas system based on Pleistocene


refuges. In M. S. Milano, L. Y. Takahashi, & M. L. Nunes (Eds.), Unidades de conservação:
Atualidades e tendencias (pp. 76–91). Fundaçao o Boticario.

Page 48 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Wetterberg, G. B., Pádua, M. T. J., Castro, C. S., & Vasconcelos, J. M. C. (1977). Uma
análise de prioridades em conservação da natureza na Amazônia. PNUD/FAO/IBDF/BRA.
45 IBDF, Brasilia, DF Série Técnica No. 8.

White, A., Khare, A., & Motnar, A. (2005). Who owns, who conserves and why it mat­
ters. Forest Trends, Washington, DC.

Whitmore, T. C. (1978). Gaps in the forest canopy. In P. B. Tomlinson & M. H. Zimmerman


(Eds.), Tropical trees as living systems (pp. 639–655). Cambridge, U.K.: Cambridge Uni­
versity Press.

Whittaker, R. H. (1962). Classification of natural communities. Botanical Review, 28(1), 1–


239.

Wilcox, B. A. (1980). Insular ecology and conservation. In M. E. Soulé & B. A. Wilcox


(Eds.), Conservation biology (pp. 95–117). Sunderland, MA: Sinauer.

Wilkie D. S., Bennett, E. L., Peres, C. A., & Cunningham, A. A. (2011). The year in ecolo­
gy and conservation biology: The empty forest revisited. Annals of the New York
Academy of Sciences, 1223(2011), 120–128.

Woodwell, G. (1978). The carbon dioxide question. Scientific American, 238(1), 34–43.

Woodwell, G., Hobbie, J. E., Houghton, R. A., Melillo, J. M., Moore, B., Peterson, B. J., &
Shaver, G. R. (1983). Global deforestation: Contribution to atmospheric carbon dioxide.
Science, 222, 1081–1086.

World Commission on Environment and Development. (1987). Our common future. Annex
to General Assembly Document A/42/427, Development and international co-operation:
Environment, August 2.

World Wildlife Fund. (2009). Amazon alive! A decade of discovery 1999–2009. Brasília,
Brazil: World Wildlife Fund.

Young, C. E. F., & Medeiros, R. (Eds.). (2018). Quanto vale o verde: A importância
econômica das unidades de conservação brasileiras. Rio de Janeiro, Brazil: Conservation
International.

Zahawi, R. A., Holl, K. D., Cole, R. J., & Reid, J. L. (2013). Testing applied nucleation as a
strategy to facilitate tropical forest recovery. Journal of Applied Ecology, 50, 88–96.

Zamora, C. (1971). Capacidad de uso de los suelos del Perú (3ª aprox.). Lima, Peru: Ofici­
na Nacional de Evaluación de Recursos naturales, ONERN.

Notes:

(1.) RAISG. 1997–2019. La Red Amazónica de Información Socioambiental Georreferenci­


ada. Provides multiple periodically updated information on the Amazon situation.

Page 49 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020


Conservation in the Amazon: Evolution and Situation

Marc Dourojeanni

National Agrarian University, La Molina

Page 50 of 50

PRINTED FROM the OXFORD RESEARCH ENCYCLOPEDIA,  ENVIRONMENTAL SCIENCE (oxfordre.com/environmen­


talscience) (c) Oxford University Press USA, 2020. All Rights Reserved. Personal use only; commercial use is strictly prohibit­
ed (for details see Privacy Policy and Legal Notice).

date: 30 April 2020

You might also like