You are on page 1of 88

251

ISSN 0041-6436

An international journal of forestry and forest industries Vol. 70 2019/1

FORESTS: NATURE-BASED
SOLUTIONS FOR WATER
Forest and Water Programme
The FAO Forest and Water Programme this vision by facilitating the sharing
envisions a world in which resilient of knowledge and experiences,
forest landscapes are managed developing the capacity of forest,
effectively to provide sustainable land and water managers to manage
water ecosystem services. In the forest–water nexus, and
collaboration with partners from the providing tools to support
forest and water sectors, it supports decision-making.
countries and stakeholders in realizing

More information:
www.fao.org/in-action/forest-and-
water-programme

Unasylva is published in English, French and Spanish. Subscriptions can be obtained by Cover: Aerial view of a village near Phang
sending an e-mail to unasylva@fao.org. Subscription requests from institutions (e.g. libraries, Nga Bay, Thailand. Forests and water have
companies, organizations and universities) rather than individuals are preferred in order to always been inextricably entwined.
make the journal accessible to more readers. All issues of Unasylva are available online free of © iStock.com/Oleh Slobodeniuk
charge at www.fao.org/forestry/unasylva.
Comments and queries are welcome at unasylva@fao.org
251ISSN 0041-6436

An international journal of forestry and forest industries Vol. 70 2019/1

Editor: A. Sarre
Editorial Advisory Board: N. Berrahmouni, J.
Campbell, P. Csoka, J. Fox, H. Abdel Hamied, D.
Contents
Hewitt, T. Hofer, H. Ortiz, L. Pina, E. Springgay,
A. Taber, S. Wertz, Xia, Z., E. Yazici, Zhang, D. Editorial 2
Emeritus Advisers: J. Ball, I.J. Bourke,
C. Palmberg-Lerche, L. Russo E. Springgay
Proofreader: Jana Gough Forests as nature-based solutions for water 3
Designer: Roberto Cenciarelli
D. Ellison, L. Wang-Erlandsson, R. van der Ent and M. van Noordwijk
The designations employed and the presentation of material Upwind forests: managing moisture recycling for
in this information product do not imply the expression of any
opinion whatsoever on the part of the Food and Agriculture nature-based resilience 14
Organization of the United Nations (FAO) concerning the legal
or development status of any country, territory, city or area or A.D. del Campo, M. González-Sanchis, U. Ilstedt, A. Bargués-Tobella
of its authorities, or concerning the delimitation of its frontiers
or boundaries. The mention of specific companies or products and S. Ferraz
of manufacturers, whether or not these have been patented,
does not imply that these have been endorsed or recommended
Dryland forests and agrosilvopastoral systems: water at the core 27
by FAO in preference to others of a similar nature that are not
mentioned. M. Gustafsson, I. Creed, J. Dalton, T. Gartner, N. Matthews, J. Reed,
The views expressed in this information product are those of L. Samuelson, E. Springgay and A. Tengberg
the author(s) and do not necessarily reflect the views or policies Gaps in science, policy and practice in the forest–water nexus 36
of FAO.

ISBN 978-92-5-131910-9 R. Lindsay, A. Ifo, L. Cole, L. Montanarella and M. Nuutinen


© FAO, 2019 Peatlands: the challenge of mapping the world’s invisible stores
of carbon and water 46
D.W. Hallema, A.M. Kinoshita, D.A. Martin, F.-N. Robinne, M. Galleguillos,
Some rights reserved. This work is made available under S.G. McNulty, G. Sun, K.K. Singh, R.S. Mordecai and P.F. Moore
the Creative Commons Attribution-NonCommercial-
ShareAlike 3.0 IGO licence (CC BY-NC-SA 3.0 IGO; https:// Fire, forests and city water supplies 58
creativecommons.org/licenses/by-nc-sa/3.0/igo/legalcode).
L. Spurrier, A. Van Breda, S. Martin, R. Bartlett and K. Newman
Under the terms of this licence, this work may be copied,
redistributed and adapted for non-commercial purposes, Nature-based solutions for water-related disasters 67
provided that the work is appropriately cited. In any use of
this work, there should be no suggestion that FAO endorses
any specific organization, products or services. The use of the
FAO Forestry 75
FAO logo is not permitted. If the work is adapted, then it must
be licensed under the same or equivalent Creative Commons World of Forestry 76
licence. If a translation of this work is created, it must include
the following disclaimer along with the required citation:
“This translation was not created by the Food and Agriculture
Books 78
Organization of the United Nations (FAO). FAO is not
responsible for the content or accuracy of this translation. The
original [Language] edition shall be the authoritative edition.”

Disputes arising under the licence that cannot be settled


amicably will be resolved by mediation and arbitration as
described in Article 8 of the licence except as otherwise
provided herein. The applicable mediation rules will be
the mediation rules of the World Intellectual Property
Organization http://www.wipo.int/amc/en/mediation/rules
and any arbitration will be conducted in accordance with
the Arbitration Rules of the United Nations Commission on
International Trade Law (UNCITRAL).

Third-party materials. Users wishing to reuse material


from this work that is attributed to a third party, such as
tables, figures or images, are responsible for determining
whether permission is needed for that reuse and for obtaining
permission from the copyright holder. The risk of claims
resulting from infringement of any third-party-owned
component in the work rests solely with the user.

Sales, rights and licensing. FAO information products are


available on the FAO website (www.fao.org/publications)
and can be purchased through publications-sales@fao.org.
Requests for commercial use should be submitted via: www.
fao.org/contact-us/licence-request. Queries regarding rights
and licensing should be submitted to: copyright@fao.org.
EDITORIAL

W
ater – clean, drinkable water – is likely to be one of relationship between forests and water) in policies and practice.
the most limiting resources in the future, given the Managing this nexus will be crucial for achieving many of the
growing global population, the high water demand Sustainable Development Goals, but it requires taking a landscape
of agricultural production systems and urban centres, and the approach. The ability to do this suffers from a lack of knowledge
confounding effects of climate change. We need to manage water about the factors that regulate the multiple functions of landscapes,
wisely – efficiently, cost-effectively and equitably – if we are to their interactions, and, ultimately, their effects on water users.
avoid the calamity of a lack of usable water. The authors describe opportunities to address the forest–water
Forested watersheds provide an estimated 75 percent of the nexus at the landscape scale, and they make recommendations
world’s accessible freshwater resources, on which more than half for research to help fill the gaps in knowledge.
the Earth’s people depend for domestic, agricultural, industrial Lindsay et al. make the case for much more policy attention on
and environmental purposes. Sustainable forest management is peatlands, which, they say, are often unrecognized or ignored and
essential, therefore, for good water management, and it can pro- therefore subject to widespread drainage and land-use conversion.
vide “nature-based solutions” for many water-related challenges. Yet peatlands contain huge stores of carbon and their destruction
This edition of Unasylva explores the challenges in realizing or mismanagement, therefore, could add substantially to global
the potential. warming. For example, even a shallow peat (30 cm deep) contains
In her article, Springgay explains that nature-based solutions more carbon than does primary tropical rainforest. Peatlands are
in water management involve the management of ecosystems also huge freshwater reservoirs and their loss could have major
(forested or otherwise) to mimic or optimize natural processes in implications for the sustainability of water supplies. Part of the
the provision and regulation of water. In many parts of the world problem in gaining more recognition for peatlands is that they
today, water management relies largely on “grey” infrastructure can be difficult to identify, and the authors provide a simple test;
involving the use of concrete and steel. A move towards nature- they also make recommendations for policymakers on how to
based solutions, says Springgay, requires a transformative shift in tackle this substantial but largely hidden challenge.
thinking in which forests and other ecosystems are viewed and Hallema and co-authors look at the implications of chang-
managed as freshwater regulators. She makes several recommen- ing forest fire regimes for forest and water management. The
dations to facilitate the transition towards “green” infrastructure increasing occurrence of extreme wildfires is threatening the
in water management. capacity of forests to deliver clean water. The authors say that
In their article, Ellison et al. present startling findings on the developing cost-effective strategies for managing fire and water
role of forests in multiplying the oceanic supply of freshwater in light of climate change, increasing urbanization and other
through moisture recycling (in which rainfall is returned to the trends requires a better understanding of the regional impacts and
atmosphere through evapotranspiration, making it available interactions of fire. Forests that are important for water supply
downwind to fall again as rain). Forests, say the authors, exhibit but at risk of extreme wildfire need to be identified and actively
more intense moisture recycling than non-forest land cover, managed, requiring the involvement of forest managers, hydrolo-
partly because of their larger water-storage potential, which, in gists, wildfire scientists, public-health specialists and the public.
turn, enables them to return rainfall to the atmosphere even in Finally, Spurrier et al. look at the crucial role of mangroves in
dry periods. Mapping the sources and sinks of precipitation and reducing the risk of disaster for millions of vulnerable coastal
evaporation can indicate where forest restoration efforts will people. Despite their importance, mangroves continue to decline
be most effective in maximizing moisture recycling for drier in extent, and climate change and other pressures threaten them
areas downwind. There is a desperate need, say the authors, to further. To help maintain the disaster-risk-reduction role of man-
redesign institutional frameworks to take into account long- groves and other natural (or green) infrastructure, the authors
distance forest–water relationships and their feedback effects recommend the use of adaptive frameworks and decision-support
on water availability. tools that enable managers to integrate and continuously update
Del Campo and co-authors present three case studies to show projections of climate-change risk, land use and human population
how “water-centred” management approaches can increase the growth.
resilience of dryland forests in the face of climate change. For Forests and water have always been inextricably entwined,
example, judicious management of Aleppo pine forest in a dry and forest managers have always needed to consider hydrology
region of Spain can increase tree growth and vigour and protect in their management decisions. But as resources become more
soils while adding to catchment water budgets and downstream constrained and water demand grows ever greater, water manage-
water flows. Such “ecohydrological-based forest management” ment will inevitably come even more to the fore in forest-related
can increase water availability in water-limited environments decision-making. Recognizing the importance of the forest–water
and therefore also socio-ecological resilience. nexus is the first step in building it into institutional processes
Gustaffson and co-authors look at gaps in the knowledge and finding forest-based solutions for water.
required to fully incorporate the forest–water nexus (i.e. the
3
© FAO/DANIEL HAYDUK

Forests as nature-based solutions for water


E. Springgay

G
A transformation is needed from rowing populations and increasing approaches to water management are
conventional forest management industrialization, urban develop- inadequate for ensuring the well-being
approaches to nature-based ment and demand for food and of human populations, biodiversity and
solutions that make water-related consumer goods have led to large-scale ecosystems.
ecosystem services the primary land-cover and land-use change globally, An estimated 65 percent of water fall-
objective. which has, in turn, caused hydrological ing on land is either stored within soil or
changes. It is also increasingly apparent evaporated from soil and plants (Oki and
that much of the human-made grey-water Kanae, 2006), with 95 percent of the soil
infrastructure,1 such as dams, pipes, ditches water stored within or above groundwater
and pumps, has contributed to global zones (Bockheim and Gennadiyev, 2010).
pro­blems and that business-as-usual Therefore, terrestrial ecosystems are
important for land–water–energy balances,
influencing soil water and atmospheric
Grey infrastructure generally refers to engi-
1

neering projects that use concrete and steel, moisture availability and thus affecting
green infrastructure depends on plants and eco­
systems, and blue infrastructure combines green
Elaine Springgay is Forestry Officer at the FAO spaces with good water management (Sonneveld Above: Forests as a nature-based solution
Forestry Department, Rome, Italy. et al., 2018). for water, United Republic of Tanzania

Unasylva 251, Vol. 70, 2019/1


4

climate (Huntington, 2006; Ellison et al., to achieve the social, economic and water for human consumption, industry
2017; Creed and van Noordwijk, 2018). All environmental goals embedded in these; and the environment.
forests influence water (FAO, 2018b), from it is essential, therefore, to strategically Land-use decisions can have significant
cloud forests and tree-covered wetlands integrate natural solutions, including consequences for water resources, com-
upstream to dryland and coastal forests green and blue infrastructure, into overall munities, economies and environments
downstream. It has been estimated that management approaches. The integration in distant (downstream and downwind)
forested watersheds provide 75 percent of of nature-based solutions shows promise locations. The loss of natural forests may
the world’s accessible freshwater resources for addressing water scarcity through increase water yields in the short term
and that more than half the Earth’s popula- supply-side management, particularly by but have long-term negative impacts on
tion is dependent on these water resources increasing water quality and groundwater water quantity and quality. For example,
for domestic, agricultural, industrial and recharge, which ultimately is essential for evapotranspiration from the Amazon River
environmental purposes (Millennium sustainable food production, improved and Congo River basins is a major source
Ecosystem Assessment, 2005). Forests are human settlements, access to water supply of precipitation (around 50–70 percent)
sometimes referred to as natural infrastruc- and sanitation, water-related risk reduc- in the Rio de la Plata basin and the Sahel,
ture, and their management can provide tion, and building resilience to climate respectively (Van der Ent et al., 2010;
“nature-based solutions” for a range of variability and change (UNWWDR, 2018). Ellison et al., 2017). Large-scale forest
water-related societal challenges. This It is estimated that USD 10 trillion will loss and land conversion affect these
article explores that potential. need to be invested in grey infrastructure natural processes, reducing cloud cover
between 2013 and 2030 for adequate and precipitation downwind (Ellison et
FORESTS: NATURAL water management (Dobbs et al., 2013). al., 2017; Creed and van Noordwijk, 2018).
INFRASTRUCTURE FOR WATER Nature-based solutions could reduce this Forest restoration and tree planting will
Nature-based solutions are actions that investment burden while also improving likely improve water quality, with the
protect, sustainably manage and restore economic, social and environmental out- impacts of such interventions depend-
natural and modified ecosystems in ways comes. Nearly USD 24 billion is estimated ing on species, management regime and
that effectively and adaptively address to have been spent on green infrastructure temporal and spatial scale. It is estimated
societal challenges and deliver benefits for water in 2015, benefiting 487 million that land conservation and restoration,
for human well-being and biodiversity hectares of land (Bennet and Ruef, 2016). including forest protection, reforestation
(Cohen-Shacham et al., 2016). In water Paying greater attention to landscape man- and agroforestry, could lead to a reduc-
management, nature-based solutions agement, including integrated watershed tion of 10 percent or more in sediments
involve the management of ecosystems management, land protection, reforestation and nutrients in watersheds (Abell et al.,
to mimic or optimize natural processes, and riparian restoration, could reduce the 2017). Care is needed, however, to ensure
such as vegetation, soils, wetlands, water operational and maintenance costs of that achieving water-quality goals does
bodies and even groundwater aquifers, grey infrastructure (Echavarria et al., not result in unacceptable trade-offs with
for the provision and regulation of water. 2015; Box 1). water yield.
The adoption of nature-based solutions In addition to their water-related eco­
for water requires a transformative shift in The role of forests in hydrology system services, forests provide habitat for
thinking from demand- to supply-oriented All forests affect hydrology and so, there- fish and other aquatic species, which, in
water management and planning, in which fore, does their management. Forests turn, play roles in ensuring the function-
crucial ecosystems such as forests are seen and trees use water and provide many ality of these ecosystems. The quantity,
not only as users but also as regulators of provisioning, regulating, supporting and quality, temperature and connectivity of
fresh water. cultural ecosystem services. Forested water resources influence fish populations
Nature-based solutions have gained areas and landscapes with trees, there- and aquatic biodiversity. Changes in these
attention in recent years because of their fore, are integral components of the water factors can affect species richness, even-
potential for addressing water scarcity cycle, regulating streamflow, fostering ness and endemism, thus influencing the
and contributing to the achievement of the groundwater recharge and contributing biodiversity and food systems of dependent
Sustainable Development Goals (SDGs), to atmospheric water recycling, including populations.
the Paris Agreement on climate change, cloud generation and precipitation through Many fish and other aquatic organisms
the Sendai Framework for Disaster Risk evapotranspiration. Forested areas and are sensitive to ecosystem degradation,
Reduction, the Aichi Biodiversity Targets, landscapes with trees also act as natural such as through eutrophication, habitat
and other international commitments. Grey filters, reducing soil erosion and water degradation and fragmentation, acidifi-
infrastructure alone will be insufficient sedimentation, thus providing high-quality cation, and changes in temperature and

Unasylva 251, Vol. 70, 2019/1


5

Box 1
Forest management: nature-based solution for urban water supply

Ninety percent of major cities rely on forested watersheds for their water supply (McDonald and Shemie, 2014), with one-third of the world’s
largest cities, including Bogotá, Johannesburg, New York, Tokyo and Vienna, obtaining a large proportion of their drinking water from pro-
tected forest areas (Dudley and Stolton, 2003).
Source-water protection, including through forest restoration and trees on agricultural land, could improve water quality for more than 1.7
billion people living in cities at a cost of less than USD 2 per person per year (which would be offset by savings from reduced water treatment)
(World Bank, 2012; Abell et al., 2017). For example, a forest-based initiative to reduce water pollution from agriculture has saved the City
of New York from the need to install a treatment plant (at an estimated cost of USD 8 billion–10 billion), as well as an additional USD 300
million per year in operational and maintenance costs. New York City has the largest unfiltered water supply in the United States of America
(Abell et al., 2017). Similarly, the estimated water conservation value of Beijing’s forests is USD 632 million (approximately USD 689 per
ha) per year (Biao et al., 2010).
Forests are used as nature-based solutions for water-related natural hazards. In Peru’s Pacific Coast water basin, where an estimated two-thirds
of historical tree cover has been lost (WRI, 2017), integrating green and grey infrastructure could reduce Lima’s dry-season deficit by 90
percent, and this would be more cost-effective than implementing grey infrastructure alone (Gammie and de Bievre, 2015). Likewise, local
forest restoration is being used in Malaga, Spain, to mitigate flood risk.
As urban populations grow, ecosystems and their services will increasingly be pushed to their limits (Kalantari et al., 2018). This is particularly
true in the fastest-growing cities – small and medium-sized cities that are undergoing rapid and mostly unplanned expansions of their urban
areas but which may need to rely increasingly on watersheds for water supply. Of the three fastest-growing cities in Africa and Asia (based
on United Nations data), an unpublished FAO review has determined that only Kampala, Uganda, acknowledges the water-related services
provided by forests.
The potential of forest management to provide nature-based solutions to mitigate some of the challenges of urban development needs to be
considered in spatial planning and management strategies (Kalantari et al., 2018). To grow sustainably, cities will need to play active roles in
protecting the water sources on which they depend.

Children cross
a river in the
Philippines. It
is important to
manage forests
© FAO/JAKE SALVADOR

and trees with


water ecosystem
services in mind
and to maximize
the forest benefits
for water

Unasylva 251, Vol. 70, 2019/1


6

climate (FAO, 2018a). For example, the six in the past 100 years, in direct cor- health and well-being, the environment
number of threatened and endangered relation with population growth (Wada et and sustainable development (Veolia and
freshwater species has increased due to the al., 2016); water consumption continues IFPRI, 2015). For example, an estimated
poor health of inland water systems (FAO, to grow at about 1 percent per year (FAO, 80 percent of all industrial and municipal
2018a). The Living Planet Index indicates undated). The global population is pro- wastewater is released into the environment
an 83 percent decline in freshwater species jected to increase from 7.7 billion in 2017 to without treatment (WWAP, 2017). Changes
populations since 1970 (WWF, 2018). 9.4 billion–10.2 billion people in 2050, with in water sediment loads and temperature
Forests and trees can help mitigate minor two-thirds living in cities (United Nations, can significantly affect fish populations
to moderate flooding events, control ava- 2018). Global water demand is projected and aquatic biodiversity, which may fur-
lanches, combat desertification, and abate to rise by 20–30 percent by 2050, due to ther affect dependent food chains and food
storm surges. For example, mangrove for- population growth, associated economic security (FAO, 2018a).
ests act as protective shields against wind development, changing consumption pat- Changes in land cover and use, population
and wave erosion, storm surges and other terns, land-use change and climate change, growth, and the frequency and intensity of
coastal hazards (FAO, 2007; Nagabhatla, among other factors (Burek et al., 2016; extreme events associated with a changing
Springgay and Dudley, 2018), and trees in WWAP, 2018). climate increase the risk of water-related
drylands help abate soil erosion and drought Under a business-as-usual scenario, the disasters. Since 1992, floods, droughts and
by capturing fog water, reducing surface world is projected to face a 40 percent water storms have affected 4.2 billion people and
water runoff and promoting groundwater deficit by 2050 (WWAP, 2015). Domestic caused USD 1.3 trillion in damage world-
recharge (Ellison et al., 2017). Changes in water use will increase significantly in wide (UNESCAP/UNISDR, 2012). Floods
land use – such as large-scale deforesta- all regions, particularly Africa and Asia, have become more frequent, increasing
tion or, conversely, forest restoration – can where domestic demand is expected to from an average of 127 events per year
influence the resilience of landscapes in triple, and in Central and South America, in 1995–2004 to 171 events per year in
the face of water-related natural hazards. where estimated future demand is double 2005–2014; floods have accounted for 47
It is important, therefore, to manage current withdrawals (Burek et al., 2016). At percent of all weather-related disasters
forests and trees with water ecosystem the same time, food demand is expected to since 1995 and affected 2.3 billion people
services in mind and to maximize the forest increase by 60 percent, requiring more land (CRED and UNISDR, 2015).
benefits for water and mitigate negative for food production and causing impacts It is estimated that floods, droughts and
impacts. A range of management deci- on soil and water resources that likely will storms result in average global losses
sions, such as species selection, stocking lead to further degradation (FAO, 2011b). of USD 86 billion per year across all
densities, and location in the landscape, Meanwhile, less than 1 percent of the economic sectors, with Africa and Asia
can have important effects on hydrology. total available freshwater is allocated for most affected in terms of deaths, dam-
Managing forests for multiple benefits is maintaining the health of ecosystems that aged communities and economic losses.
the foundation of sustainable forest man- serve as natural infrastructure for water The cost of floods, droughts and storms
agement, but it requires an understanding (Boberg, 2005; Nagabhatla, Springgay and worldwide is expected to escalate to USD
and recognition of trade-offs. For example, Dudley, 2018). 200 billion–400 billion per year by 2030
fast-growing exotic species planted for bio- Approximately 80 percent of the world’s (OECD, 2015).
mass and carbon sequestration may have a population suffers from moderate to severe The impacts of disasters could be miti-
positive impact on water quality but could water scarcity (Mekonnen and Hoekstra, gated if land and forest conversion, urban
greatly reduce water supply. Reducing tree 2016). Nearly half the global population is expansion and planning, and the intensifica-
densities, prolonging rotation cycles and already living in areas with potential water tion of food production take ecological
conserving native forests in riparian buffer scarcity at least one month per year, and it functions into account and aim to improve
zones could mitigate these negative effects. is estimated that this will increase to 4.8 – rather than degrade – ecosystem services.
billion–5.7 billion people – more than half 2
According to the Ramsar Convention on Wet-
WATER: A GLOBAL CHALLENGE the projected global population – by 2050 lands (2016), wetlands “are areas of marsh, fen,
Davidson (2014) estimated that up to 87 (Burek et al., 2016). peatland or water, whether natural or artificial,
percent of all wetlands, 2 including tree- Water pollution has worsened in almost permanent or temporary, with water that is static
or flowing, fresh, brackish or salt, including areas
covered wetlands and peatlands, have been all rivers in Africa, Asia and Latin of marine water the depth of which at low tide
lost worldwide since the eighteenth cen- America since the 1990s (UNEP, 2016; does not exceed six metres”. They “may also
tury; up to 71 percent of all wetlands have WWAP, 2018), and the degradation of incorporate riparian and coastal zones adjacent
to the wetlands, and islands or bodies of marine
been destroyed since 1900. Global water water resources is expected to increase water deeper than six metres at low tide lying
consumption has increased by a factor of in the next decades, threatening human within the wetlands”.

Unasylva 251, Vol. 70, 2019/1


7

© FAO/JOAN MANUAL BALIELLAS


A forest park in Hanoi,
Viet Nam. Forests and
water go arm-in-arm

Box 2
Incentivizing forest–water management
Payment for ecosystem services (PES) schemes constitute a potential incentive mechanism for better environmental management. Applied to
forest–water management, PES schemes require service “buyers” (usually downstream communities and industries) and service “providers”
(upstream communities who are considered forest stewards). PES schemes have limitations, however: for example, they rely on the complex
valuation of ecosystem services, often require formal land-tenure arrangements, depend on evidence that services are delivered, and can have
implications for socio-economic power dynamics. These limitations may explain the lack of successful PES schemes.
Other incentive mechanisms exist. For example, “reciprocal watershed agreements” are simple grassroots versions of conditional transfers
that help land managers in upper watershed areas to sustainably manage their forest and water resources in ways that benefit both themselves
and downstream water users. Like PES, reciprocal watershed agreements depend on an understanding that hydrological services are being
provided, and they rely on recognized conditions of tenure at the local level (i.e. who owns, controls and grants access to watershed forests).
In contrast to PES schemes, however, reciprocal watershed agreements offer demand-led rewards rather than monetary incentives, with
compensation based on specific needs that diversify income sources. For example, downstream water users could provide upstream landowners
with improved livelihood options such as beehives, fruit-tree seedlings and better irrigation equipment (Porras and Asquith, 2018).
Reciprocal watershed agreements have been implemented successfully in Bolivia (Plurinational State of), where more than 270 000 water users
have signed agreements with 6 871 upstream landowners to conserve 367 148 ha of water-producing forests. In return, the reciprocity-based
conservation agreements provide sufficient funding for alternative development projects such as drip irrigation, fruit and honey production and
improved cattle management. Fifty-two municipalities in the country have adopted such agreements since 2003 (Natura Foundation, 2019).
The success of reciprocal watershed agreements in Bolivia (Plurinational State of) may be due partly to the fact that the agreements have
been made in areas with cloud forests: people can see that deforestation reduces dry-season flows and that improved cattle management that
restricts livestock movement improves water quality. In such cases, upstream conservation measures can easily be shown to contribute to the
protection of watershed services – without the need for detailed and costly hydrological assessments.
In addition, scale and local perceptions of forest–water links matter. The watersheds subject to the agreements are small, and there is a limited
number of land uses and stakeholders; it is easier, therefore, to see the benefits of improved management, and land managers and water users
can easily be identified. Moreover, the mechanism is likely to be more successful in areas where local stakeholders already understand and
perceive the links between forest management and maintaining healthy freshwater ecosystems.

Unasylva 251, Vol. 70, 2019/1


8

A GLOBAL PICTURE OF FORESTS Despite growing recognition of the for example, the United States Forest
AND WATER influence and importance of forests for Service identifies itself as the manager of
An estimated 31 percent of the global land water, only 25 percent of forests globally the nation’s largest water resource (United
area is forested, of which 65 percent is are managed with soil and water conser- States Forest Service, 2017). Europe falls
degraded (FAO, 2010; 2015). The World vation as one of the primary objectives below the global average of managing
Resources Institute calculates tree-cover (Figure 1). Moreover, a little less than 10 forests for soil and water conservation
trends by major water basin, 3 or hydro- percent of forests is managed primarily because most forest land is privately owned
shed, as well as water-related hazard risk for soil and water conservation, includ- and is not accounted for in national report-
(i.e. erosion, forest fire and baseline water ing around 2 percent managed primarily ing; however, a recent report provided
stress4). Before 2000, hydrosheds averaged for clean water and about 1 percent each ample evidence of integrated approaches to
68 percent tree cover; this had reduced to for coastal stabilization and soil erosion forest–water management in Europe (FAO
31 percent by 2000, however, and to 29 control (FAO, 2015). Only 13 countries and UNECE, 2018). In many countries in
percent by 2015. This tree-cover loss has report that all their forests are managed the tropics and subtropics, however, there
not necessarily been evenly distributed: with consideration given to soil and water
approximately 38 percent of the hydro- conservation. 3
FAO divides the world into 230 major basins or
sheds had lost more than half their tree More than 70 percent of forests in North watersheds (FAO, 2011a).
cover by 2000, rising to 40 percent by 2014 America are managed with consider- 4
Baseline water stress is defined as the ratio of
total water withdrawals to total renewable water
(WRI, 2017). ations for soil and water conservation; supply in a given area (WRI, 2017).
1
Percentage area of
forests for soil and
water conservation
by country and
forest type

0
<10
10−25
25−50
50−75
75−90
>90
no data

1990 2000 2005 2010 2015

500 000
HECTARES

400 000

300 000

200 000

100 000

0
BOREAL TEMPERATE SUBTROPICAL TROPICAL

Source: FAO (2018b).

Unasylva 251, Vol. 70, 2019/1


9

is a negative trend in the area of forests as floods, forest fires, landslides and storm Sustainable Development Goals
managed for soil and water conservation, surges. Of the hydrosheds that had lost The interconnection between forests and
and deforestation is also ongoing. Although at least half their tree cover by 2015, 88 water is explicitly referenced in two SDGs:
all forests have impacts on hydrology, the percent had a medium to very high risk of SDG 6 (“Clean water and sanitation”) and
loss of tropical and subtropical forests erosion, 68 percent had a medium to very SDG 15 (“Life on land”). In SDG target
may disproportionately affect the global high risk of forest fire, and 48 percent had a 6.6, forests are recognized as water-related
hydrological cycle (FAO, 2018b). medium to very high risk of baseline water ecosystems; similarly, SDG target 15.1
Decreases in tree cover can lead to stress (WRI, 2017) (Figure 2). refers to forests as freshwater ecosystems.
increased soil erosion and degradation Although the indicators for these targets
and, in turn, to a reduction in water qual- BUILDING ON INTERNATIONAL do not measure the interlinkages between
ity. In some cases, the loss of tree cover MOMENTUM forests and water, methodologies exist
is also associated with reduced water The notion of forest management as a for looking at this relationship – which
availability, especially when natural for- nature-based solution for water is not new. countries could use to better understand
ests are converted to other land uses that The forest–water relationship is a cross- and report on how forests serve as natu-
degrade or compact soils, thus reducing cutting issue, and it has gained increased ral infrastructure for water. For example,
soil infiltration, water storage capacity and attention in the last two decades (Figure 3). in addition to the indicator used in the
groundwater recharge (Bruijnzeel, 2014; The UN Decade on Ecosystem Restoration FAO Global Forest Resources Assessment
Ellison et al., 2017; FAO, 2018b). The forest (2021–2030) will undoubtedly raise the (“area of forests managed for soil and water
loss and degradation associated with land profile of forest management as a nature- conservation”), Ramsar (2019) specifies
conversion and poor land management based solution for water to new heights other forested or tree-covered areas, such
practices may also increase the risk to and because of the wide-ranging potential as peatlands, as wetlands. Around 123
damage from water-related hazards, such impacts of restoration on hydrology and million ha of forest – about 2.9 percent of
the need to take these into account in plan- the world’s forest area – are classified as
2 ning restoration initiatives. Ramsar sites.
Hydroshed risk to erosion and base water
stress, by percentage tree cover loss

Erosion risk by % tree cover loss Baseline water stress risk Forest fire risk by
by % tree cover loss (2015) % tree cover loss (2015)

35 35 35

30 30 30

Very high Very high Very high


High High High
Medium 25 Medium 25 Medium 25
Low Low Low
Very low Very low Very low

20 20 20

15 15 15

10 10 10

5 5 5

0 0 0
90% 80% 70% 60% 50% 40% 30% 20% 10% 90% 80% 70% 60% 50% 40% 30% 20% 10%
90% 80% 70% 60% 50% 40% 30% 20% 10%

Source: WRI (2017).

Unasylva 251, Vol. 70, 2019/1


10

3
Recent global forest policy initiatives that could encourage nature-based solutions for water

Sustainable Development Goals


• SDG 6: Clean water and sanitation
• SDG 13: Climate action
• SDG 14: Life below water
• SDG 15: Life on land
• Other SDGs also apply, including SDG 1 (No poverty); SDG 2 (Zero hunger); SDG 8 (Decent work and economic
growth); and SDG 11 (Sustainable cities and communities)

United Nations Convention to Combat Desertification


• Strategic Objective 1: To improve the condition of affected ecosystems, combat desertification/land degradation,
promote sustainable land management and contribute to land degradation neutrality
• Strategic Objective 2: To improve the living conditions of affected populations
• Strategic Objective 3: To mitigate, adapt to and manage the effects of drought in order to enhance the resilience
of vulnerable populations and ecosystems
• Strategic Objective 4: To generate global environmental benefits through effective implementation of the Convention

Convention on Biological Diversity Aichi Targets


• Target 1: People are aware of the values of biodiversity and the steps they can take to conserve and use it sustainably
• Target 4: Sustainable production and consumption with impacts of use of natural resources well within safe
ecological limits
• Target 5: Rate of loss of all natural habitats, including forests, is at least halved and where feasible brought close
to zero, and degradation and fragmentation is significantly reduced
• Target 7: Areas under agriculture, aquaculture and forestry are managed sustainably, ensuring conservation of
biodiversity
• Target 11: Terrestrial and inland water, and coastal and marine areas of particular importance for biodiversity and
ecosystem services, are conserved
• Target 14: Ecosystems that provide essential services, including services related to water, and contribute to health,
livelihoods and well-being, are restored and safeguarded
• Target 15: Ecosystem resilience and the contribution of biodiversity to carbon stocks has been enhanced through
conservation and restoration

Other international processes


• United Nations Framework Convention on Climate Change – countries have made commitments under the Paris
Agreement through their nationally determined contributions and national adaptation plans
• Sendai Framework for Disaster Risk Reduction: Priority 1 – Understanding disaster risk; Priority 2 – Strengthen-
ing disaster risk governance to manage disaster risk; Priority 4 – Enhancing disaster preparedness for effective
response and to “build back better” in recovery, rehabilitation and reconstruction
• Ramsar Convention on Wetlands: Strategic Goal 1 – Addressing the drivers of wetland loss and degradation;
Strategic Goal 3 – Wisely using all wetlands
• Forest landscape restoration – countries have made commitments on land restoration by 2030, many including
water-related objectives

Unasylva 251, Vol. 70, 2019/1


11

(Intended) nationally determined To do this, a scientific understanding of the and building social and environmental
contributions context is needed, including the well-being resilience (FAO and UNECE, 2018).
Forests and water resources feature and needs of communities and ecosystems. A key challenge of management is to
prominently in the nationally determined A transformation in approach may be optimize the multiple benefits and mini-
contributions of countries to the Paris required for a rapid transition from tradi- mize the trade-offs.
Agreement on climate change. Eighty- tional forest management options, such as • Increase connectivity within
eight percent of the original “intended” silviculture for wood production or conser- and between landscapes. Hydrol-
nationally determined contributions of vation, to regimes in which the provision of ogy connects landscapes, including
countries referenced forests as part of land water-related ecosystem services is the pri- upstream and downstream water bod-
use, land-use change and forestry, and 77 mary objective. Nature-based solutions do ies and related terrestrial ecosystems;
percent referenced water (French Water not necessarily require additional financial atmospheric water teleconnects land-
Partnership and Coalition Eau, 2016). resources; rather, they have the potential to scapes at the continental scale. The
Forty-nine percent of 168 (intended) enable the more effective use of existing conservation and restoration of upland
nationally determined contributions refer financing (WWAP, 2018) by increasing the forests and peatlands, the establishment
to the interlinkages between forest and value of multiple forest goods and services, of riparian networks, and the restora-
water management, including references including water, and reducing investments tion of meandering water courses and
to integrated (water) resource manage- in grey infrastructure. wetlands will help maintain the hydro-
ment and the water ecosystem services The following recommendations are logical functionality of landscapes, and
provided by forests and mangroves, with made to facilitate the rapid transition restored areas will also function as bio-
the majority of these references included towards nature-based solutions for water. diversity corridors for terrestrial and
under adaptation measures. Of the coun- • Implement science-based manage- aquatic species.
tries indicating their nationally determined ment and guidelines. Forest man- • Greatly intensify collaboration
contributions, those in Africa, Asia and agement for water ecosystem services among sectors. The integration
Latin America give most recognition to not only needs to take into account of nat u r a l a nd hu m a n-m a d e
the importance of forest management as a current environmental and socio- infrastructure is needed to address
nature-based solution for water (Springgay economic conditions but also future global water, land and urban
et al., forthcoming). projections related to land-use planning challenges effectively. This requires
Although nationally determined and climate scenarios. The aim of spe- forestry to collaborate with other
contributions do not imply resource cies selection, spacing, thinning and sectors, including water, agriculture,
commitments until 2020, the strong rotation cycles should be to optimize urban planning, disaster risk man-
acknowledgement of forest–water rela- water ecosystem services, biomass and agement and energy. Collaboration
tionships within them suggests there carbon storage and manage potential between ministries in governments
is significant political will to address trade-offs. Examples exist of landscape poses well-known challenges; at the
the issue, offering an opportunity to management (such as ecosystem-based local level, on the other hand, many
promote the integration of forests as management) that prioritizes eco­system stakeholders – governments, landown-
nat u r a l i n f r a st r uct u r e i n wat er integrity and functionality, and these ers and businesses – are involved in
management. could be more widely employed and multiple sectors as managers of lands
integrated. and forests and their associated water
GLOBAL CHALLENGES NEED • Bundle benefits in schemes to better resources. Is it possible to engage with
CROSS-CUTTING, INTEGRATED compensate landowners and man- other sectors without fighting for juris-
SOLUTIONS agers for their water management dictional control? The forest sector
Changing the landscape changes the practices. Managing forests for water should consider marketing its skills
hydrology. This is true for all scenarios – can produce a wide range of other in forest management and long-term
whether tree-cover loss results in land-use goods and services, including carbon planning to other sectors reliant on
change, or a degraded landscape is restored sequestration, biodiversity conserva- sustainable forest and tree manage-
through reforestation or afforestation. To tion, cultural services (e.g. education ment as a nature-based solution for
fully take into account the impacts of and recreation), and wood and non- the immense challenges facing our
forest-related landscape change on water wood forest products. The bundling water resources. u
in land management decisions, it is neces- of the multiple benefits of forests is
sary to consider temporal and spatial scales a cost-effective means for increasing
as well as short- and long-term objectives. income opportunities for communities

Unasylva 251, Vol. 70, 2019/1


12

Centre for Research on the Epidemiology of FAO. 2018a. Review of the State of the World
Disasters (CRED) and United Nations Office Fishery Resources: inland fisheries. FAO
for Disaster Risk Reduction (UNISDR). Fisheries and Aquaculture Circular No. C942
References Creed, I.F. & van Noordwijk, M., eds. 2018. Rev.3. Rome.
Forest and water on a changing planet: FAO. 2018b. The State of the World’s Forests
Abell, R., Asquith, N., Boccaletti, G., Bremer, vulnerability, adaptation and governance 2018: Forest pathways to sustainable
L., Chapin, E., Erickson-Quiroz, A. & opportunities. A global assessment report. development. Rome (available at www.fao.
Higgins, J., et al. 2017. Beyond the source: IUFRO World Series, Volume 38. Vienna, org/3/I9535EN/i9535en.pdf).
the environmental, economic and community International Union of Forest Research FAO. Undated. AQUASTAT [online]. FAO.
benefits of source water protection. Organizations (IUFRO). [Cited June 2019]. fao.org/nr/water/aquastat/
Arlington, USA, The Nature Conservancy. Davidson, N.C. 2014. How much wetland main/index.stm
Bennett, G. & Ruef, F. 2016. Alliances for has the world lost? Long-term and FAO & UNECE. 2018. Forests and water:
Green Infrastructure: State of Watershed recent trends in global wetland area. valuation and payments for forest ecosystem
Investment 2016. Washington DC, Forest Marine and Freshwater Research, 65(10): services (available at www.unece.org/
Trends Ecosystem Marketplace (available 934–941. fileadmin/DAM/timber/publications/sp-44-
at www.forest-trends.org/documents/files/ Dobbs, R., Pohl, H., Lin, D., Mischke, forests-water-web.pdf).
doc_5463.pdf). J., Garemo, N., Hexter, J., Matzinger, French Water Partnership & Coalition
Biao, Z., Wenhua, L., Gaodi, X. & Yu, X. 2010. S., Palter, R. & Nanavatty, R. 2013. Eau. 2016. Review of the integration of
Water conservation of forest ecosystem in Infrastructure productivity: how to save water within the Intended Nationally
Beijing and its value. Ecological Economics, $1 trillion a year. McKinsey Global Institute. Determined Contributions (INDCs) for
69(7): 1416–1426. Dudley, N. & Stolton, S. 2003. Running pure: COP21 (available at www.coalition-eau.
Boberg, J. 2005. Freshwater availability. In: the importance of forest protected areas org/wpcontent/uploads/2016-03-review-
J. Boberg, Liquid assets: how demographic to drinking water. A research report for of-water-integration-in-indc-vf.pdf).
changes and water management policies the World Bank and WWF Alliance for Gammie, G. & de Bievre, B. 2015. Assessing
affect freshwater resources, pp. 15–28. Forest Conservation and Sustainable Use. green interventions for the water supply of
Santa Monica, Arlington & Pittsburgh, USA, Washington, DC, World Bank. Lima, Peru. Report for Forest Trends and
RAND Corporation. Echavarria, M., Zavala, P., Coronel, L., CONDESAN. Washington, DC.
Bockheim, J.G. & Gennadiyev, A.N. 2010. Montalvo, T. & Aguirre, L.M. 2015. Green Huntington, T.G. 2006. Evidence for
Soil-factorial models and earth-system infrastructure in the drinking water sector in intensification of the global water cycle:
science: a review. Geoderma, 159(3–4): Latin America and the Caribbean: trends, review and synthesis. Journal of Hydrology,
243–51. challenges, and opportunities. 319(1–4): 83–95.
Bruijnzeel, L.A. 2014. Hydrological functions Ellison, D., Morris, C.E., Locatelli, B.D., Kalantari, Z., Ferreira, C., Keesstra, S.
of tropical forests: not seeing the soil for Sheil, D., Cohen, J., Murdiyarso, D. & & Destouni, G. 2018. Ecosystem-based
the trees? Agriculture, Ecosystems and Gutierrez, V. 2017. Trees, forests and solutions for flood-drought risk mitigation in
Environment, 104: 185–228. water: cool insights for a hot world. Global vulnerable urbanizing parts of East Africa.
Burek, P., Satoh, Y., Fischer, G., Kahil, M.T., Environmental Change, 43: 51–61. Current Opinion in Environmental Science
Scherzer, A., Tramberend, S., Nava, L.F., FAO. 2007. Mangroves of Asia 1980–2005. & Health, 5: 73–78.
Wada, Y., Eisner, S., Flörke, M., Hanasaki, Countr y reports. Forest Resources McDonald, R.I. & Shemie, D. 2014. Urban
N., Magnuszewski, P., Cosgrove, B. Assessment Programme Working Paper water blueprint: mapping conservation
& Wiberg, D. 2016. Water futures and No. 137. Rome. solutions to the global water challenge
solution: fast track initiative (final report). FAO. 2010. Global Forest Resources [online]. The Nature Conservancy. [Cited
IIASA Working Paper. Laxenburg, Austria, Assessment 2010: Main report. FAO June 2019]. http://water.nature.org/
International Institute for Applied Systems Forestry Paper No. 163. Rome. waterblueprint/#/intro=true
Analysis (IIASA). FAO. 2011a. World map of the major Mekonnen, M.M. & Hoekstra, A. 2016. Four
Cohen-Shacham, E., Walters, G., Janzen, hyd rological basins (der ived f rom billion people facing severe water scarcity.
C. & Maginnis, S., eds. 2016. Nature- HydroSHEDS). Science Advances, 2(2).
based solutions to address global societal FAO. 2011b. Save and grow: a policy maker’s Millennium Ecosystem Assessment. 2005.
challenges. Gland, Switzerland, International guide to the sustainable intensification of Global assessment report. Washington, DC,
Union for Conservation of Nature. the smallholder crop production. Rome. Island Press.
CRED & UNISDR. 2015. The human cost FAO. 2015. Global Forest Resources Nagabhatla, N., Springgay, E. & Dudley, N.
of weather-related disasters 1995–2015. Assessment 2015. Rome. 2018. Forests as nature-based solutions for

Unasylva 251, Vol. 70, 2019/1


13

ensuring urban water security. Unasylva, UNEP. 2016. A snapshot of the world’s water fs.fed.us/science-technology/water-air-soil
250: 43–52. quality: towards a global assessment. Van der Ent, R.J., Savenije, H.H.G., Schaefli,
Natura Foundation. 2019. Watershared Nairobi, United Nations Environment B. & Steele-Dunne, S.C. 2010. Origin and
Program [online]. [Cited June 2019]. www. Programme (UNEP). fate of atmospheric moisture over continents.
naturabolivia.org/en/watershared-2 UNESCAP/ UNISDR. 2012. Reducing Water Resources Research, 49(9).
OECD. 2015. OECD Principles on water vulnerability and exposure to disasters. Veolia & IFPRI. 2015. The murky future of
governance [online]. Organisation for The Asia-Pacific Disaster Report 2012. global water quality: new global study
Economic Co-operation and Development Bangkok, United Nations Economic and projects rapid deterioration in water quality.
(OECD). [Cited June 2019]. www.oecd. Social Commission for Asia and the Pacific Washington, DC, International Food Policy
org/governance/oecd-principles-on-water- (UNESCAP)/United Nations Office for Research Institute (IFPRI)/Veolia.
governance.htm Disaster Risk Reduction (UNISDR). Wada, Y., Flörke, M., Hanasaki, N., Eisner,
Oki, T. & Kanae, S. 2006. Global hydrological United Nations. 2018. 2018 Revision of world S., Fischer, G., Tramberend, S., Satoh, Y.,
cycles and world water resources. Science, urbanization prospects [online]. [Cited Van Vliet, M.T.H., Yillia, P., Ringler, C.,
313(5790): 1068–1072. June 2019]. www.un.org/development/ Burek, P. & Wiberg, D. 2016. Modelling
Porras, I. &. Asquith, A. 2018. Ecosystems, desa/publications/2018-revision-of-world- global water use for the 21st century: the
povert y alleviation and conditional urbanization-prospects.html Water Futures and Solutions (WFaS)
transfers. London, International Institute United Nations World Water Assessment initiative and its approaches. Geoscientific
for Environment and Development (IIED). Programme (WWAP). 2015. The United Model Development, 9: 175–222.
Ramsar Convention Secretariat. 2016. An Nations World Water Development Report World Bank. 2012. Inclusive green growth:
introduction to the Ramsar Convention on 2015: Water for a sustainable world. Paris, the pathway to sustainable development.
Wetlands. Gland, Switzerland (available United Nations Educational, Scientific and Washington, DC.
at www.ramsar.org/sites/default/files/ Cultural Organization (UNESCO). WRI. 2017. Global Forest Water Watch [online].
documents/libra r y/ handbook1_ 5ed _ United Nations World Water Assessment World Resource Institute (WRI). [Cited June
introductiontoconvention_e.pdf Programme (WWAP). 2017. The United 2019]. www.globalforestwatch.org.
Ramsar Sites Information Service. 2019. Nations World Water Development Report WWF. 2018. Living Planet Report 2018:
Ramsar Sites Information Service [online]. 2017: Wastewater: the untapped resource. Aiming higher. M. Grooten & R.E.A.
[Cited June 2019]. https://rsis.ramsar.org Paris, United Nations Educational, Scientific Almost, eds. Gland, Switzerland, WWF. u
Sonneveld, B.G.J.S., Merbis, M.D., Alfarra, and Cultural Organization (UNESCO).
A., Ünver, O. & Arnal, M.A. 2018. Nature- United Nations World Water Assessment
based solutions for agricultural water Programme (WWAP). 2018. The United
management and food security. FAO Land Nations World Water Development Report
and Water Discussion Paper no. 12. Rome, 2018: Nature-based solutions for water.
FAO. Licence: CC BY-NC-SA 3.0 IGO. Paris, United Nations Educational, Scientific
Springgay, E., Casallas, S., Janzen, S. & and Cultural Organization (UNESCO).
Brito, V.V. Forthcoming. The forest–water United States Forest Service. 2017. Water, air,
nexus: an international perspective. Forests. and soil [online]. [Cited June 2019]. www.

Unasylva 251, Vol. 70, 2019/1


14

Upwind forests: managing moisture recycling for


nature-based resilience
D. Ellison, L. Wang-Erlandsson, R. van der Ent and M. van Noordwijk

E
Trees and forests multiply the fficient and effective forest and however, tends to focus on river flows and
oceanic supply of freshwater water-related nature-based solu- to take rainfall for granted as an unruly,
through moisture recycling, tions to challenges in human devel- unmanageable input to the system (Ellison,
pointing to an urgent need to halt opment require a holistic understanding Futter and Bishop, 2012). Thus, the poten-
deforestation and offering a way to of the role of forest–water interactions tial impact of increased tree and forest
increase the water-related benefits in hydrologic flows and water supply in cover on downwind rainfall and potential
of forest restoration. local, regional and continental landscapes. water supply is both underestimated and
Forest and water resource management, underappreciated.

Afternoon clouds over the Amazon rainforest

© NASA IMAGE COURTESY JEFF SCHMALTZ, MODIS RAPID RESPONSE AT NASA GSFC

David Ellison is at the Department of Forest


Resource Management, Swedish University of
Agricultural Sciences, Umeå, Sweden, Adjunct
Researcher, Sustainable Land Management
Unit, Institute of Geography, University of Bern,
Switzerland, and at Ellison Consulting, Baar,
Switzerland.
Lan Wang-Erlandsson is at the Stockholm
Resilience Centre, Stockholm University,
Stockholm, Sweden.
Ruud van der Ent is at the Department of Water
Management, Faculty of Civil Engineering and
Geosciences, Delft University of Technology,
Delft, the Netherlands, and the Department of
Physical Geography, Faculty of Geosciences,
Utrecht University, Utrecht, the Netherlands.
Meine van Noordwijk is at the World
Agroforestry Centre, Bogor, Indonesia, and Plant
Production Systems, Wageningen University,
Wageningen, the Netherlands.

Unasylva 251, Vol. 70, 2019/1


15

On average, about 60 percent of all 10 percent of the Earth’s land surface but al., 2014, 2010; Gebrehiwot et al., 2019).
transpiration and other sources of ter- contribute 22 percent of global evapotrans- The long-distance relationships between
restrial evaporation (jointly referred to as piration (Wang-Erlandsson et al., 2014), forests, moisture recycling and rainfall
evapotranspiration) returns as precipita- an important share of which returns to challenge conventional forest–water
tion over land through terrestrial moisture land as rainfall. Moreover, deep-rooted analyses based on catchments as the
recycling, and approximately 40 percent trees are able to access soil moisture and principal unit of analysis (Ellison, Futter
of all terrestrial rainfall originates from groundwater and thus continue to tran- and Bishop, 2012; Wang-Erlandsson et al.,
evapotranspiration (van der Ent et al., 2010; spire during dry periods when grasses are 2018). Catchment-centric studies tend to
see also Figure 1). From the perspective of dormant, providing crucial moisture for ignore evapotranspiration once it has left
a river, evapotranspiration may appear as rainfall when water is most scarce (Staal the confines of the basin in which it was
a loss but, for the extended landscape, the et al., 2018; Teuling et al., 2010). produced, despite its key contributions
recycling of atmospheric moisture (“rivers Nature-based solutions involving for- elsewhere to downwind rainfall (Ellison,
in the sky”) supports downwind rainfall. est and landscape restoration, therefore, Futter and Bishop, 2012) – and the view
Forests are disproportionately impor- have the potential to influence rainfall that evapotranspiration represents a loss
tant for rainfall generation. On average, and consequently sometimes very dis- rather than a contribution to the hydrologic
their water use is 10–30 percent closer tant, downwind rainfall systems reliant cycle has resulted in a pronounced bias
to the climatically determined potential on moisture recycling for food produc- both against forests and in favour of the
evapotranspiration than that of agricul- tion, water supply and landscape resilience catchment-based water balance (Bennett
tural crops or pastures (Creed and van (Bagley et al., 2012; Dirmeyer et al., and Barton, 2018; Dennedy-Frank and
Noordwijk, 2018). For example, tropical 2014; Dirmeyer, Brubaker and DelSole,
evergreen broadleaf forests occupy about 2009; Ellison et al., 2017; van der Ent et 1
The global hydrologic landscape

ATMOSPHERIC MOISTURE TRANSPORT


45 FO
HUMIDITY AND
PRECIPITATION BIOPRECIPITATION
RECYCLING TRIGGERS
LOCAL AND GLOBAL DOWNWIND
AT REGIONAL
75 E L
120 P L

SCALE HEATING AND CONTINENTAL


AND COOLING SCALE

FOG/CLOUD
455 E O
410 P O

INTERCEPTION
FLOOD
MODERATION

DOWNSTREAM INFILTRATION
AND GROUNDWATER
RECHARGE
5

OCEAN LAND
45 FL

Notes: F represents “net” atmospheric moisture exchange between land (L) and ocean (O). Inflows of atmospheric moisture to land from the ocean are, on average, about
75 000 km3 per year, significantly larger than the “net” inflows of 45 000 km3 suggest (van der Ent et al., 2010). Likewise, the evapotranspiration contribution to rainfall over
oceans is approximately 30 000 km3 per year (van der Ent et al., 2010).
Sources: Adapted from Ellison et al. (2017), with quantifications of water flow (i.e. ocean evaporation, EO; evapotranspiration, EL; ocean precipitation, PO; land precipitation,
PL; net ocean-to-land moisture flow, FO, rainbow arrow; and runoff, FL, black arrow) in 1 000 km3 per year from van der Ent and Tuinenburg (2017).

Unasylva 251, Vol. 70, 2019/1


16

Gorelick, 2019; Filoso et al., 2017; Jackson atmospheric long-distance forest–water form of vapour and drops (i.e. evapotrans-
et al., 2005; Trabucco et al., 2008). relationships, and discuss some of the key piration and precipitation); and, second,
New modelling capacities and increased challenges and opportunities for using for- those that flow horizontally as atmospheric
data availability, however, make it pos- ests as nature-based solutions for water. moisture (thus, rivers in the sky) (Figure 1).
sible for scientists to better and more easily Our focus is on the role of forests for On average, approximately 75 000 km3 of
quantify where and how much forests rainfall and water supply through mois- water per year evapotranspires from land
contribute to rainfall. The last decade ture recycling. Thus, we ignore the many into the atmosphere, where it combines
has seen a surge, not only in understand- other invaluable benefits of forest–water with evaporation of oceanic origin (Oki
ing of the forest–rainfall relationship interactions, such as flood moderation, and Kanae, 2006; Rodell et al., 2015;
through moisture recycling, but also in water purification, infiltration, groundwater Trenberth, Fasullo and Mackaro, 2011).
the scientific exploration of landscape, recharge and terrestrial surface cooling (see Of the evapotranspiration from land, some
forest and water management and gov- Ellison et al., 2017). falls as rain over oceans, but 60 percent –
ernance opportunities (Creed and van about 45 000 km3 per year – falls as rainfall
Noordwijk, 2018; Ellison et al., 2017; Keys FORESTS SUPPLY AND MULTIPLY over land (Dirmeyer et al., 2014; van der
et al., 2017). FRESHWATER RESOURCES Ent et al., 2010). In total, evapotranspira-
In this article we review the role of The global distribution of moisture tion contributes approximately 40 percent
forests as water recycler and water-resource recycling of the 120 000 km3 of water per year that
multiplier, examine the implications of The largest water flows over land are not precipitates over land.
those in rivers but rather those that “invisi- Trees, forests and other vegetation
Trees contribute to evapotranspiration bly” flow first in the vertical direction in the play pivotal roles in supporting both
by accessing deep soil moisture and
groundwater, as well as through interception
© JOAKIM WANG-ERLANDSSON

Unasylva 251, Vol. 70, 2019/1


17

evapotranspiration and precipitation. On Gordon, 2001), as opposed to evaporation under current atmospheric circulation con-
a global average, transpiration makes from bare soil or open water evaporation ditions. In large parts of Europe, the eastern
up about 60 percent of total evapotrans­ (Miralles et al., 2016; Wang-Erlandsson Russian Federation, East Africa and north-
piration, with a large uncertainty range et al., 2014). Climate model simulations ern South America, more than one-third of
(Coenders-Gerrits et al., 2014; Schlesinger suggest that a green planet with maximum evapotranspiration is vegetation-regulated
and Jasechko, 2014; Wang-Erlandsson et vegetation could supply three times as (i.e. occurs because of the presence of veg-
al., 2014; Wei et al., 2017). Vegetation’s much evapotranspiration from land and etation) and falls as precipitation over land
direct contribution to total evapotrans- twice as much rainfall as a desert world (Figure 3, p. 21). In parts of Eurasia,
piration, however, also includes canopy, with no vegetation (Kleidon, Fraedrich North America, southern South America
forest-floor and soil-surface evaporation, as and Heimann, 2000). and large parts of subtropical and dryland
well as epiphyte interception. Significantly Tree-, forest- and vegetation-regulated Africa, more than one-third of precipitation
more than 90 percent of total terrestrial moisture recycling is unevenly distributed. comes from vapour flows that would not
evapotranspiration comes from vegetated Figure 2a shows the rainfall-generation ben- occur without vegetation (Keys, Wang-
land (Abbott et al., 2019; Rockström and efits provided by existing vegetation cover Erlandsson and Gordon, 2016).

2
a) Share of
evapotranspiration
that is vegetation-
regulated and falls
as precipitation over
land (%)

b) Share of
precipitation that
comes from upwind
vegetation-regulated
evapotranspiration (%)

Notes: The figure shows the relative importance of current global vegetation for evaporation that returns as precipitation on land (top panel), and
precipitation that originates as evapotranspiration on land (bottom). The estimates are based on model coupling between the hydrologic model
STEAM and the moisture-tracking model WAM-2layers, simulating a “current land” and a “barren land/sparse vegetation” scenario. “Vegetation-
regulated” evapotranspiration and precipitation are defined as the difference in evapotranspiration and precipitation between these two scenarios.
The destination of evapotranspiration and origin of precipitation are subsequently determined using WAM-2layers. These model simulations
capture the immediate interactions with the atmospheric water cycle but do not consider changes in circulation, soil quality, runoff and water
availability.
Source: Keys, Wang-Erlandsson and Gordon (2016), used here under a CC BY 4.0 licence.

Unasylva 251, Vol. 70, 2019/1


18
© LAN WANG-ERLANDSSON

Trees contribute to the redistribution of both


stream and atmospheric moisture flows
Key aspects of forest moisture recy- roughly half the evapotranspiration returns
cling: moisture retention and rainfall as rainfall over land (van der Ent et al.,
Most regions of the world are essentially multiplier 2010). Where rainfall exceeds actual
dependent, to varying degrees, on the abil- In general, heavily forested regions amounts of evapotranspiration, rivers are
ity of landscapes to recycle moisture to exhibit more intense moisture recycling fed by surplus flows. Thus, where forest
downwind locations. Without vegetation- than non-forested regions. During wet loss breaks the moisture recycling chain,
regulated precipitation, a significant share periods, transpiration, rainfall and the there are potentially cascading downwind
of rainfall across land surfaces would be water intercepted by leaves in a forest are consequences for both rainfall and river
lost. Moreover, vegetation regulation can closely related to each other in time and flows (Ellison et al., 2017; Gebrehiwot
critically influence the length of grow- space. The average distance that water et al., 2019; Lovejoy and Nobre, 2018;
ing seasons and becomes even more particles travel from forested regions Molina et al., 2019; Nobre, 2014; Sheil
important in dry periods (Keys, Wang- during the wet season can be as low as and Murdiyarso, 2009; Wang-Erlandsson
Erlandsson and Gordon, 2016). Thus, 500–1 000 km, especially in rainforest (van et al., 2018).
considerable benefit can be obtained der Ent and Savenije, 2011). Evaporated Further, forests differ crucially from
from restoring very large shares of moisture from denser rainforests spends shorter vegetation types in their larger
deforested and degraded landscapes (on average) less than five days in the water-storage potential – below the ground,
with trees and forests in order to sustain atmosphere (van der Ent and Tuinenburg, on the forest floor and in the canopy. This
and intensify the hydrologic cycle and 2017). This illustrates the ability of for- storage allows trees to return significantly
thus increase the availability of fresh- ests to create their own rainfall. In large more rainfall to the atmosphere as evapo-
water resources on terrestrial surfaces. parts of the Amazon and Congo basins, transpiration over longer periods of time,

Unasylva 251, Vol. 70, 2019/1


19

even without rain. Soil-moisture storage, and green water,1 where the total amount from clouds and fog represent additional
therefore, enables forests to play an espe- of water available is influenced solely by benefits from adding tree and forest cover
cially important role in the water cycle interannual climatic variation in the total (Bright et al., 2017; Bruijnzeel, Mulligan
when water is most scarce. Forests develop quantities of precipitation. Based on this and Scatena, 2011; Ellison et al., 2017;
deep roots to cope with droughts, in con- newer understanding of the hydrologic Ghazoul and Sheil, 2010; Hesslerová et
trast to shorter vegetation types, which cycle, rainfall is an endogenous systemic al., 2013).
tend to go dormant (Wang-Erlandsson et element and responds to changing land-use
al., 2016). With deeper roots, trees are able conditions within and across landscapes. NATURE-BASED SOLUTIONS AND
to both store and access more water in ECOSYSTEM-BASED ADAPTATION
the soil, which they use for transpiration Moisture recycling and the role of To facilitate a moisture-recycling-
during periods without rain (Teuling et al., catchments based rethinking of trees and forests
2010) as well as to tap into groundwater For the most part, moisture recycling makes as nature-based solutions, we highlight
resources (Fan et al., 2017; Sheil, 2014). its principal contributions at distances well key differences in the consideration of
This transpired moisture generates dry- beyond the catchment scale. This can pres- green- and blue-water availability; the
season rainfall in more-distant regions (van ent a dilemma for local water-resource multiple benefits of forest-supplied mois-
der Ent et al., 2014), which can be essen- managers because planting more trees and ture recycling; the precipitationshed and
tial for buffering ecosystems, farmlands forests in an individual catchment will evaporationshed as conceptual tools; and
and human communities against drought typically have the effect of flushing more challenges for the governance of forest-
(Staal et al., 2018). Because dry seasons water resources out of the same catch- moisture recycling across competing
and droughts often mean declines in the ment and into the atmosphere (Bennett and interests and scales.
supply of ocean evaporation to land, the Barton, 2018; Calder et al., 2007; Dennedy-
relative role of forests can be heightened Frank and Gorelick, 2019; Filoso et al., Rethinking total available water:
in dry periods (Bagley et al., 2012). The 2017; Jackson et al., 2005). Where the the difference between green
ability of forests to retain moisture and locally available water supply is limited, and blue water
release it in dry periods can help stabilize reforestation may need to be undertaken From the catchment perspective, it may
and extend growing seasons – which may in other upwind locations or atmospheric appear to make sense to start from mea-
be especially crucial in places experiencing outflows from the catchment compensated. sured precipitation as the expression
a climate-change-induced increase in dry Locally, this can be achieved by reducing of total available water supply (Gleick
spells and dry seasons. other catchment-based water uses, such as and Palaniappan, 2010; Hoekstra and
The ability of forests to retain and pro- those involving croplands, industries and Mekonnen, 2012; Mekonnen and Hoekstra,
vide moisture for multiple cycles of rainfall human populations. Regionally, reforesta- 2016; Schyns et al., 2019; Schyns, Booij
recycling means that forests not only “re- tion efforts may need to be coordinated so and Hoekstra, 2017). This would ignore,
allocate” a fixed amount of precipitation that increased evapotranspiration-related however, evapotranspiration – the “green”
but also both multiply that amount and catchment outflows are compensated by production of atmospheric moisture – by
further alter the temporal dynamics of increased precipitation inflows from addi- trees, forests, croplands and other forms
precipitation. This perspective contrasts tional upwind reforestation. of vegetation (van Noordwijk and Ellison,
sharply with conventional catchment-based Not all catchments are water-challenged, 2019). Through moisture recycling, vegeta-
water resource management, which consid- and many can benefit from additional forest tion makes water from upwind oceanic
ers the total amount of water available on restoration. Thus, in water-rich and flood- sources available across ever more distant
terrestrial surfaces as a fixed quantity in prone catchments, trees and forests can inland locations and regulates the climate
a zero-sum allocation game between blue aid the redistribution of water resources by cooling terrestrial surfaces (Bagley et
to downwind communities while simul- al., 2012; Ellison et al., 2017; Ellison,
The green and blue water paradigm divides
1
taneously facilitating local infiltration, Futter and Bishop, 2012; van der Ent et al.,
up the catchment water balance into multiple
components. Green water represents all water soil storage and groundwater recharge 2010; Keys, Wang-Erlandsson and Gordon,
that is evapotranspired back to the atmosphere (Bargués Tobella et al., 2014; Bruijnzeel, 2016; van Noordwijk et al., 2014; Sheil
by trees, plants, croplands and open water bodies. 2004; Ilstedt et al., 2016; McDonnell et al., and Murdiyarso, 2009; Wang-Erlandsson
Blue water represents the remaining surface
and groundwater that is available for human 2018). Moreover, adding more trees and et al., 2018).
consumption and industrial use. Grey water, forests can help moderate flooding (van Along upwind coasts, the appropriation
generally not discussed here, represents water Noordwijk, Tanika and Lusiana, 2017) and of one unit of freshwater for human or
that has been degraded through industrial or
human use (Falkenmark and Rockström, 2006; reduce erosion. The cooling of terrestrial industrial consumption is worth many
Hoekstra, 2011). surfaces and the absorption of moisture times the same amount in downwind

Unasylva 251, Vol. 70, 2019/1


20
ALEXIS BROSS IS LICENSED UNDER CC BY-NC-SA 2.0 

Deforestation-induced reductions in
rainfall not only affect ecosystems it might be more useful to consider “poten- “maximum vegetation” scenario (i.e. 100
and agriculture but also the water
tially available” water. This can largely percent dense forest cover over land) could
supplies of cities, such as the
megacity of Tokyo, Japan be considered a function of three factors: be almost twice that of a desert world, or
1) how much of the upwind local catchment about 137 000 km3 of precipitation per
water availability. Thus, different ele- water balance can be recycled back into year compared with 71 000 km3 per year
ments of the blue, green and grey water the atmosphere for potential downwind in the “no-vegetation” scenario, due to
paradigm cannot be treated as removable rainfall; 2) how many times the oceanic increased water recycling and surface
or interchangeable modular units that contribution to the terrestrial water budget radiation and despite increased cloud cover.
can simply be plugged into or out of a can be recycled in this way; and 3) the Their estimate suggests a doubling of the
system at will. The whole is not equal to extent to which increased recycling can evapotranspiration-to-land precipitation
the sum of its parts (van Noordwijk and dampen dry spells and shorten the length ratio relative to a desert world and sug-
Ellison, 2019). An alternative – but rarely of dry seasons. gests a potential addition of some 17 000
recognized – strategy for managing and Given that 40–50 percent of the world’s km3 in total annual rainfall compared to
potentially improving catchment-based forests have already been removed from the current total annual rainfall estimated
water availability is therefore to increase terrestrial surfaces (Crowther et al., 2015), in Figure 1. 2 In less-extreme scenarios
the amount of upwind forest cover in order a crucial question is: How much additional and assuming fixed moisture-recycling
to bring more rainfall to downwind basins freshwater could be added to the terrestrial
(Creed and van Noordwijk, 2018; Dalton et water budget by progressively restoring 2
Global hydrologic cycle estimates of total annual
al., 2016; Ellison, 2018; Keys et al., 2012; previously forested and currently degraded rainfall vary in the range of approximately
99 000–129 000 km 3 (Abbott et al., 2019;
Weng et al., 2019). landscapes? The extreme-scenario simula- Trenberth et al., 2011). Thus, incorporating this
In contrast to the predom inant tion by Kleidon, Fraedrich and Heimann uncertainty into the estimate by Kleidon, Frae-
catchment-centric approach to measuring (2000), based on one climate model, sug- drich and Heimann (2000) yields an approximate
range of +8 000–+37 000 km3 per year.
and allocating terrestrial water resources, gested that terrestrial precipitation in a

Unasylva 251, Vol. 70, 2019/1


21

ratios, another study suggested that poten- solutions. Payment schemes for ecosys- thereby reducing evapotranspiration and
tial vegetation (i.e. the natural potential tem services (Martin-Ortega, Ojea and downwind precipitation. Decreased pre-
vegetation state under current climate Roux, 2013) are a possible means by which cipitation, in turn, increases the risk of fire
conditions) could lead to an additional 600 such strategies could be implemented on (IUFRO, 2018), which can cause forest loss
km3 of terrestrial precipitation per year the ground. To date, however, we are or even self-amplified forest dieback (Staal
compared with current land use (Wang- unaware of any ecosystem-based adap- et al., 2015; Zemp et al., 2017). Because
Erlandsson et al., 2018). This scenario tation efforts aimed explicitly at putting of the large carbon stores, rich biodiver-
includes irrigation, which provides higher moisture‑recycling principles into practice sity and climate regulation provided by
evapotranspiration and precipitation than (Creed and van Noordwijk, 2018), despite tropical forests, forest dieback risks trig-
“potential vegetation”. the great potential of such forest and land- gering further climate change, cascading
In both estimates, the accumulated global scape restoration strategies. On the other regime shifts and teleconnected circulation
increase in potential precipitation and hand, models are being developed for when shifts (Boers et al., 2017; Lawrence and
water availability masks important spatial and where additional reforestation could be Vandecar, 2015; Rocha et al., 2018).
heterogeneity. Large uncertainties around considered to increase moisture recycling Agriculture is not only a major driver
the effects of reforestation and afforesta- (Creed and van Noordwijk, 2018; Dalton of forest degradation and deforestation
tion on rainfall persist in global models et al., 2016; Ellison, 2018; Gebrehiwot (DeFries et al., 2010) but also a direct
and further analysis is needed. et al., 2019; Keys, Wang-Erlandsson and beneficiary of forest-supplied moisture.
Gordon, 2018; Wang-Erlandsson et al., Bagley et al. (2012), among others,
Nature-based solutions for whom? 2018; Weng et al., 2019). showed that crop yields in major crop-
Beneficiaries of forest-supplied rainfall Moisture recycling can have other producing regions could be affected
The role of trees and forests in maintaining important impacts on forest resilience. by land-use change through moisture
the water cycle is of broad interest and Tropical deforestation in an upwind region recycling at a magnitude similar to
points to multiple possibilities for sectoral decreases the total amount of water being climate change. Oliveira et al. (2013)
integration in the design of nature-based intercepted and stored in soil surfaces, demonstrated that agricultural expansion

3
Conceptual figure of
a precipitationshed,
in which the sink
region is selected
based (for example)
on management
interest

Source: Keys et al. (2012), used here under a CC-BY-3.0 licence.

Unasylva 251, Vol. 70, 2019/1


22

at the expense of Amazon rainforest could rainfall in other locations can be mapped restoration are likely to generate regional-
be self-defeating due to the ensuing decline in absolute (e.g. mm per year) or relative scale rainfall benefits but potentially
in rainfall. (e.g. percentage of a selected region’s decrease local river flows, local-scale
Rainfall not only feeds agriculture evaporation) terms to provide various decision-making may mis-prioritize for-
but replenishes all freshwater resources. types of information. Defining absolute est management strategies and policy.
Deforestation that reduces rainfall may precipitationshed boundaries can help This suggestion, however, runs counter
therefore also have potential consequences in identifying those regions that make to ongoing efforts in many countries to
for megacities (i.e. cities with more than the largest moisture contributions to a devolve centralized, institutional deci-
10 million inhabitants), the water supplies selected sink region’s rainfall and thus sion-making frameworks towards local
of which are taken from surface water approximately where forest protection autonomy (Creed and van Noordwijk, 2018;
(Keys, Wang-Erlandsson and Gordon, or expansion may be most advantageous Colfer and Capistrano, 2005). Striking an
2018; Wang-Erlandsson et al., 2018). For for a specific sink region. A relative appropriate balance between local gov-
example, Amazon deforestation was a precipitation­shed shows those regions ernance autonomy and the requirement
potential contributing factor in the severe with the highest contributions relative to for larger-scale water management and
2014–2017 droughts in the Brazilian mega- its own local evaporation and thus is useful for identifying and equitably sharing the
city of São Paulo (Escobar, 2015; Nazareno for screening regions where restoration cross-scale co-benefits of forest–water
and Laurance, 2015). efforts will be most cost-effective. management policies poses a considerable
challenge.
Precipitationsheds and Context-dependent governance
evaporationsheds opportunities CONCLUSION
For any area or region of interest – such Moisture-recycling governance in a given Rapidly expanding knowledge on the
as a catchment, national park, nation precipitationshed or evaporationshed is role of forest and water interactions in
or continent – the sources and sinks of highly context-dependent, varying, for moisture recycling provides important
precipitation and evaporation can be deter- example, in the number and size of the new perspectives on how trees and forests
mined through moisture tracking. As an countries involved, the heterogeneity can be used to address water scarcity in
analogue to the “watershed”, the concept of of land uses within the moisture-recy- effective nature-based solutions. Trees
the “precipitationshed” (Figure 3) defines cling domain, the nature and extent of and forests multiply the oceanic sup-
regional delineations of upwind locations regional teleconnections, and potentially ply of freshwater resources through
based on a threshold of moisture contrib- complex social dynamics (Keys et al., moisture recycling and can assist crop
uted and received (Keys et al., 2012). 2017; Keys, Wang-Erlandsson and Gordon, production by improving overall water
Studies of precipitationsheds address 2018). For example, the precipitation- availability and thereby prolonging grow-
the question: “Where does the evapora- shed of a region in Siberia (the Russian ing seasons. Without forest-supplied
tion or evapotranspiration that supplies Federation) is likely to comprise a rela- moisture, terrestrial rainfall would be
the precipitation for my selected region tively homogenous area in a single country, considerably lower in amount and extent.
occur?” The opposite question can also be whereas a similar-sized region in West Seen as an opportunity, forest-supplied
asked: “Where does the evapotranspira- Africa will encompass a wide range of moisture from upwind regions could be
tion in my selected region contribute to land-use types in several countries (Keys further enhanced by increasing forest
precipitation?” Moisture-tracking studies et al., 2017). These differences in the cover along the moisture-source trajec-
can map those areas, sometimes called specifics of particular moisture-recycling tory. In addition to enhancing moisture
evaporationsheds (e.g. van der Ent and systems are important considerations in the recycling, increasing tree and fores cover
Savenije, 2013). Watershed boundaries design of governance strategies (Keys would have other benefits for water,
are determined by landscape topography et al., 2017). such as flood moderation, water purifi-
and surface flows; precipitationsheds and Most existing transboundary water cation, increased infiltration, soil water
evaporationsheds, on the other hand, are arrangementsdo not extend beyond catch- storage, groundwater recharge and ter-
determined by atmospheric moisture flows ments or basins to include source regions restrial surface cooling.
that follow wind patterns, vary with season, of atmospheric moisture production (Creed An urgent rethinking is required of
and depend on the selection of a region of and van Noordwijk, 2018; Ellison et al., management strategies and the role of
interest for which precipitation is tracked 2017; Gebrehiwot et al., 2019; Keys et regional and national governments with
back to its evaporative source. al., 2017), despite the obvious inter- a view to creating decision-making
Both precipitationsheds and areas pro- est such arrangements should arouse. processes that can adequately consider
viding evapotranspiration that returns as Moreover, because forest protection and and better understand the current and

Unasylva 251, Vol. 70, 2019/1


23

potential future contributions of evapo- Bruijnzeel, L.A., Mulligan, M. & Scatena,


rationsheds and precipitationsheds. F.N. 2011. Hydrometeorology of tropical
Most existing forest and water manage- montane cloud forests: emerging patterns.
ment frameworks have been designed References Hydrological Processes, 25(3): 465–498.
for catch ment-cent r ic blue-water https://doi.org/10.1002/hyp.7974
upstream and downstream management. Abbott, B.W., Bishop, K., Zarnetske, J.P., Calder, I.R., Hofer, T., Vermont, S. & Warren,
But such systems entirely overlook the Minaudo, C., Chapin, F.S., Krause, S., P. 2007. Towards a new understanding of
role of moisture recycling in determining et al. 2019. Human domination of the forests and water. Unasylva, 58(229): 3–10.
the availability of freshwater resources global water cycle absent from depictions Coenders-Gerrits, A.M.J., van der Ent, R.J.,
on terrestrial surfaces. There is a desper- and perceptions. Nature Geoscience, 1. Bogaard, T.A., Wang-Erlandsson, L.,
ate need, therefore, to redesign or retrofit https://doi.org/10.1038/s41561-019-0374-y Hrachowitz, M. & Savenije, H.H.G. 2014.
existing institutional and administrative Bagley, J.E., Desai, A.R., Dirmeyer, P.A. Uncertainties in transpiration estimates.
frameworks to adequately consider long- & Foley, J.A. 2012. Effects of land cover Nature, 506(7487): E1–E2. https://doi.
distance forest–water relationships and change on moisture availability and potential org/10.1038/nature12925
their feedback effects on total water crop yield in the world’s breadbaskets. Colfer, C.J.P. & Capistrano, D., eds. 2005.
availability. Local water yields need to be Environmental Research Letters, 7(1): The politics of decentralization: forests,
considered in the context of both upwind 014009. https://doi.org/10.1088/1748- power, and people. London, UK, & Sterling,
evapotranspiration as well as downwind 9326/7/1/014009 USA, Earthscan.
contributions – that is, the regional-to- Bargués Tobella, A., Reese, H., Almaw, Creed, I.F. & van Noordwijk, M., eds. 2018.
continental-scale water balance. A., Bayala, J., Malmer, A., Laudon, Forest and water on a changing planet:
Significant and multiple benefits can be H. & Ilstedt, U. 2014. The effect of vulnerability, adaptation and governance
obtained by taking advantage of the nature- trees on preferential f low and soil opportunities. A global assessment report.
based solutions that forests can provide. infiltrability in an agroforestry parkland IUFRO World Series, Volume 38. Vienna,
Payment schemes for ecosystem ser- in semiarid Burkina Faso. Water Resources International Union of Forest Research
vices provide a potential framework Research, 50(4): 3342–3354. https://doi. Organizations (IUFRO).
for undertaking such ecosystem-based org/10.1002/2013WR015197 Crowther, T.W., Glick, H.B., Covey, K.R.,
adaptation strategies, but much more Bennett, B.M. & Barton, G.A. 2018. The Bettigole, C., Maynard, D.S., Thomas,
needs to be done to recognize and map enduring link between forest cover and S.M., et al. 2015. Mapping tree density at
out the potential. To maximize synergies, rainfall: a historical perspective on science a global scale. Nature, 525(7568): 201–205.
manage trade-offs and uncertainties, and and policy discussions. Forest Ecosystems, https://doi.org/10.1038/nature14967
overcome cross-scale ethical dilemmas, 5(1). https://doi.org/10.1186/s40663-017- Dalton, J., Ellison, D., McCartney, M.,
nature-based solutions for water 0124-9 Pittock, J. & Smith, B. 2016. Can’t see the
involving trees and forests need to be Boers, N., Marwan, N., Barbosa, H.M.J. & water for the trees? Global Water Forum, 3
co-developed in suitable institutional Kurths, J. 2017. A deforestation-induced October 2016.
arrangements that adequately recognize and tipping point for the South American DeFries, R.S., Rudel, T., Uriarte, M. &
encompass the interests of all stakeholders. monsoon system. Scientific Reports, 7: Hansen, M. 2010. Deforestation driven by
41489– 41489. https://doi.org/10.1038/ urban population growth and agricultural
ACKNOWLEDGEMENTS srep41489 trade in the twenty-first century. Nature
We are grateful to Patrick Keys for Bright, R.M., Davin, E., O’Halloran, T., Geoscience, 3(3): 178–181.
his feedback on and contributions to Pongratz, J., Zhao, K. & Cescatti, A. Dennedy-Frank, P.J. & Gorelick, S.M.
this article. Lan Wang-Erlandsson 2017. Local temperature response to land 2019. Insights from watershed simulations
acknowledges funding from the Swedish cover and management change driven by around the world: Watershed service-
Research Council Formas grant 2018- non-radiative processes. Nature Climate based restoration does not significantly
02345 Ripples of Resilience and the Change, 7(4): 296 –302. https://doi. enhance streamflow. Global Environmental
European Research Council under org/10.1038/nclimate3250 Change, 58: 101938. https://doi.org/10.1016/j.
the European Union’s Horizon 2020 Bruijnzeel, L. A. 20 04. Hydrological gloenvcha.2019.101938
research and innovation programme functions of tropical forests: not seeing the Dirmeyer, P.A., Brubaker, K.L. & DelSole,
grant agreement 743080 Earth Resilience soil for the trees? Agriculture, Ecosystems T. 2009. Import and export of atmospheric
in the Anthropocene. u & E nviron m en t, 10 4(1): 185 –228. water vapor between nations. Journal of
https://doi.org/10.1016/j.agee.2004.01.015 Hydrology, 365(1–2): 11–22. https://doi.
org/10.1016/j.jhydrol.2008.11.016

Unasylva 251, Vol. 70, 2019/1


24

Dirmeyer, P.A., Wei, J., Bosilovich, M.G. & Water, 6(1): e1317. https://doi.org/10.1002/ understand the vulnerability of rainfall
Mocko, D.M. 2014. Comparing evaporative wat2.1317 dependent regions. Biogeosciences, 9(2):
sources of terrestrial precipitation and their Ghazoul, J. & Sheil, D. 2010. Tropical rain 733–746. https://doi.org/10.5194/bg-9-733-
extremes in MERRA using relative entropy. forest ecology, diversity, and conservation. 2012
Journal of Hydrometeorology, 15(1): 102– Oxford University Press (available at Keys, P.W., Wang-Erlandsson, L. &
116. https://doi.org/10.1175/JHM-D-13-053.1 https://global.oup.com/academic/product/ Gordon, L.J. 2016. Revealing invisible
Ellison, D. 2018. From myth to concept and tropical-rain-forest-ecology-diversity-and- water: moisture recycling as an ecosystem
beyond – the biogeophysical revolution conservation-9780199285884). service. PLoS ONE, 11(3): e0151993. https://
and the forest-water paradigm. UNFF13 Gleick, P.H. & Palaniappan, M. 2010. Peak doi.org/10.1371/journal.pone.0151993
Background Analytical Study on Forests water limits to freshwater withdrawal and Keys, P.W., Wang-Erlandsson, L. & Gordon,
and Water. United Nations Forum on use. Proceedings of the National Academy L.J. 2018. Megacity precipitationsheds
Forests (UNFF) (available at http://rgdoi. of Sciences, 107(25): 11155–11162. https:// reveal tele-connected water security
net/10.13140/RG.2.2.26268.80004). doi.org/10.1073/pnas.1004812107 challenges. PLoS ONE, 13(3). https://doi.
Ellison, D., Futter, M.N. & Bishop, K. 2012. Hesslerová, P., Pokorný, J., Brom, J. & org/10.1371/journal.pone.0194311
On the forest cover–water yield debate: from Rejšková-Procházková, A. 2013. Daily Keys, P.W. , Wa ng-Erla nds son, L. ,
demand- to supply-side thinking. Global dynamics of radiation surface temperature Gordon, L.J., Galaz, V. & Ebbesson,
Change Biology, 18(3): 806–820. https:// of different land cover types in a temperate J. 2017. Approaching moisture recycling
doi.org/10.1111/j.1365-2486.2011.02589.x cultural landscape: consequences for the gover nance. Global Environmental
Ellison, D., Morris, C.E., Locatelli, B., Sheil, local climate. Ecological Engineering, Change, 45: 15–23. https://doi.org/10.1016/j.
D., Cohen, J., Murdiyarso, D., et al. 2017. 54: 145–154. https://doi.org/10.1016/j. gloenvcha.2017.04.007
Trees, forests and water: cool insights ecoleng.2013.01.036 Kleidon, A., Fraedrich, K. & Heimann,
for a hot world. Global Environmental Hoekstra, A.Y. 2011. The global dimension M. 2000. A green planet versus a desert
Change, 43: 51–61. https://doi.org/10.1016/j. of water governance: why the river basin world: estimating the maximum effect of
gloenvcha.2017.01.002 approach is no longer sufficient and why vegetation on the land surface climate.
Escobar, H. 2015. Drought triggers alarms cooperative action at global level is needed. Climatic Change, 44(4): 471–493. https://
in Brazil’s biggest metropolis. Science, Water, 3(1). https://doi.org/10.3390/w3010021 doi.org/10.1023/A:1005559518889
347(6224): 812–812. https://doi.org/10.1126/ Hoekstra, A.Y. & Mekonnen, M.M. 2012. The Lawrence, D. & Vandecar, K. 2015. Effects
science.347.6224.812 water footprint of humanity. Proceedings of of tropical deforestation on climate and
Falkenmark, M. & Rockström, J. 2006. The the National Academy of Sciences, 109(9): agriculture. Nature Climate Change,
new blue and green water paradigm: breaking 3232–3237. 5(1): 27–36. https://doi.org/10.1038/
new ground for water resources planning and Ilstedt, U., Bargués Tobella, A., Bazié, H.R., nclimate2430
management. Journal of Water Resources Bayala, J., Verbeeten, E., Nyberg, G., et al. Lovejoy, T.E. & Nobre, C. 2018. Amazon
Planning and Management, 132(3): 129– 2016. Intermediate tree cover can maximize tipping point. Science Advances, 4(2):
132. https://doi.org/10.1061/(ASCE)0733- groundwater recharge in the seasonally dry eaat2340. https://doi.org/10.1126/sciadv.
9496(2006)132:3(129) tropics. Scientific Reports, 6: 21930. aat2340
Fan, Y., Miguez-Macho, G., Jobbágy, E.G., IUFRO. 2018. Global fire challenges in a Martin-Ortega, J., Ojea, E. & Roux, C. 2013.
Jackson, R.B. & Otero-Casal, C. 2017. warming world. F-N. Robinne, J. Burns, Payments for water ecosystem services in
Hydrologic regulation of plant rooting depth. P. Kant, B. de Groot, M.D. Flannigan, M. Latin America: a literature review and
Proceedings of the National Academy of Kleine & D.M. Wotton, eds. Occasional conceptual model. Ecosystem Services,
Sciences, 114(40): 10572–10577. https://doi. Paper No. 32. Vienna, International Union 6: 122–132. https://doi.org/10.1016/j.
org/10.1073/pnas.1712381114 of Forest Research Organizations (IUFRO). ecoser.2013.09.008
Filoso, S., Bezerra, M.O., Weiss, K.C.B. Jackson, R.B., Jobbágy, E.G., Avissar, R., McDonnell, J.J., Evaristo, J., Bladon, K.D.,
& Palmer, M.A. 2017. Impacts of forest Roy, S.B., Barrett, D.J., Cook, C.W., Buttle, J., Creed, I.F., Dymond, S.F., et al.
restoration on water yield: A systematic Farley, K.A., Maitre, D.C. le, McCarl, B.A. 2018. Water sustainability and watershed
review. PloS ONE, 12(8): e0183210. https:// & Murray, B.C. 2005. Trading water for storage. Nature Sustainability, 1(8):
doi.org/10.1371/journal.pone.0183210 carbon with biological carbon sequestration. 378–379. https://doi.org/10.1038/s41893-
Gebrehiwot, S.G., Ellison, D., Bewket, W., Science, 310(5756): 1944–1947. https://doi. 018-0099-8
Seleshi, Y., Inogwabini, B.-I. & Bishop, org/10.1126/science.1119282 Mekonnen, M.M. & Hoekstra, A.Y. 2016.
K. 2019. The Nile Basin waters and the Keys, P.W., van der Ent, R.J., Gordon, L.J., Four billion people facing severe water
West African rainforest: rethinking the Hoff, H., Nikoli, R. & Savenije, H.H.G. scarcity. Science Advances, 2(2): e1500323.
boundaries. Wiley Interdisciplinary Reviews: 2012. Analyzing precipitationsheds to https://doi.org/10.1126/sciadv.1500323

Unasylva 251, Vol. 70, 2019/1


25

Miralles, D.G., Jiménez, C., Jung, M., Journal of Climate, 28(21): 8289–8318. reforestation: a global analysis of hydrologic
Michel, D., Ershadi, A., McCabe, M.F., https://doi.org/10.1175/JCLI-D-14-00555.1 impacts with four case studies. International
et al. 2016. The WACMOS-ET project – Schlesinger, W.H. & Jasechko, S. 2014. Agricultural Research and Climate
Part 2: Evaluation of global terrestrial Transpiration in the global water cycle. Change: A Focus on Tropical Systems,
evaporation data sets. Hydrology and Earth Agricultural and Forest Meteorology, 126(1–2): 81–97. https://doi.org/10.1016/j.
System Sciences, 20(2): 823–842. https://doi. 189–190: 115–117. https://doi.org/10.1016/j. agee.2008.01.015
org/10.5194/hess-20-823-2016 agrformet.2014.01.011 Trenberth, K.E., Fasullo, J. & Mackaro, J.
Molina, R.D., Salazar, J.F., Martínez, J.A., Schyns, J.F., Booij, M.J. & Hoekstra, A.Y. 2011. Atmospheric moisture transports from
Villegas, J.C. & Arias, P.A. 2019. Forest- 2017. The water footprint of wood for lumber, ocean to land and global energy flows in
induced exponential growth of precipitation pulp, paper, fuel and firewood. Advances reanalyses. Journal of Climate, 24(18): 4907–
along climatological wind streamlines in Water Resources, 107(Supplement 4924. https://doi.org/10.1175/2011JCLI4171.1
over the Amazon. Journal of Geophysical C): 490–501. https://doi.org/10.1016/j. van der Ent, R.J. & Savenije, H.H.G. 2011.
Research: Atmospheres, 124(5): 2589–2599. advwatres.2017.05.013 Length and time scales of atmospheric
https://doi.org/10.1029/2018JD029534 Schyns, J.F., Hoekstra, A.Y., Booij, M.J., moisture recycling. Atmospheric Chemistry
Nazareno, A.G. & Laurance, W.F. 2015. Hogeboom, R.J. & Mekonnen, M.M. 2019. and Physics, 11(5): 1853–1863. https://doi.
Brazil’s drought: beware deforestation. Limits to the world’s green water resources org/10.5194/acp-11-1853-2011
Science, 347(6229): 1427. https://doi. for food, feed, fiber, timber, and bioenergy. van der Ent, R.J. & Savenije, H.H.G. 2013.
org/10.1126/science.347.6229.1427-a Proceedings of the National Academy of Oceanic sources of continental precipitation
Nobre, A.D. 2014. The future climate of Sciences, 201817380. https://doi.org/10.1073/ and the correlation with sea surface
Amazonia, scientific assessment report. São pnas.1817380116 temperature. Water Resources Research,
José dos Campos, Brazil (available at www. Sheil, D. 2014. How plants water our planet: 49: 3993–4004. https://doi.org/10.1002/
ccst.inpe.br/wp-content/uploads/2014/11/ advances and imperatives. Trends in wrcr.20296
The_Future_Climate_of_Amazonia_ Plant Science, 19(4): 209–211. https://doi. van der Ent, R.J., Savenije, H.H.G., Schaefli,
Report.pdf). org/10.1016/j.tplants.2014.01.002 B. & Steele-Dunne, S.C. 2010. Origin and
Oki, T. & Kanae, S. 2006. Global hydrological Sheil, D. & Murdiyarso, D. 2009. How fate of atmospheric moisture over continents.
cycles and world water resources. Science, forests attract rain: an examination of a Water Resources Research, 46(9). https://doi.
313(579 0): 1068 –1072. ht t ps://doi. new hypothesis. BioScience, 59(4): 341–347. org/10.1029/2010WR009127
org/10.1126/science.1128845 https://doi.org/10.1525/bio.2009.59.4.12 van der Ent, R.J. & Tuinenburg, O.A.
Oliveira, L. J.C., Costa, M.H., Soares- Staal, A., Dekker, S.C., Hirota, M. & van 2017. The residence time of water in the
Filho, B.S. & Coe, M.T. 2013. Large-scale Nes, E.H. 2015. Synergistic effects of atmosphere revisited. Hydrology and Earth
expansion of agriculture in Amazonia may be drought and deforestation on the resilience System Sciences, 21(2): 779–790. https://doi.
a no-win scenario. Environmental Research of the south-eastern Amazon rainforest. org/10.5194/hess-21-779-2017
Letters, 8(2): 024021–024021. https://doi. Ecological Complexity, 22: 65–75.https:// van der Ent, R.J., Wang-Erlandsson, L.,
org/10.1088/1748-9326/8/2/024021 doi.org/10.1016/j.ecocom.2015.01.003 Keys, P.W. & Savenije, H.H.G. 2014.
Rocha, J.C., Peterson, G., Bodin, Ö. & Staal, A., Tuinenburg, O.A., Bosmans, Contrasting roles of interception and
Levin, S. 2018. Cascading regime shifts J.H.C., Holmgren, M., van Nes, E.H., transpiration in the hydrological cycle –
within and across scales. Science, 362(6421): Scheffer, M., Zemp, D.C. & Dekker, Part 2: moisture recycling. Earth Systems
1379–1383. https://doi.org/10.1126/science. S.C. 2018. Forest-rainfall cascades buffer Dynamics, 5(2): 471– 489. https://doi.
aat7850 against drought across the Amazon. Nature org/10.5194/esd-5-471-2014
Rockström, J. & Gordon, L. 2001. Assessment Climate Change, 8(6): 539–543. https://doi. van Noordwijk, M. & Ellison, D. 2019.
of green water flows to sustain major org/10.1038/s41558-018-0177-y Rainfall recycling needs to be considered
biomes of the world: implications for future Teuling, A.J., Seneviratne, S.I., Stöckli, in defining limits to the world’s green water
ecohydrological landscape management. R., Reichstein, M., Moors, E., Ciais, resources. Proceedings of the National
Physics and Chemistry of the Earth, Part P., et al. 2010. Contrasting response of Academy of Sciences, 201903554. https://
B: Hydrology, Oceans and Atmosphere, European forest and grassland energy doi.org/10.1073/pnas.1903554116
26(11–12): 843–851. https://doi.org/10.1016/ exchange to heatwaves. Nature Geoscience, va n N o o r dw ijk, M . , Na m i r e mb e,
S1464-1909(01)00096-X 3(10): 722–727. https://doi.org/10.1038/ S., Catacutan, D., Williamson, D.
Rodell, M., Beaudoing, H.K., L’Ecuyer, T.S., ngeo950 & Gebrekirstos, A. 2014. Pricing
Olson, W.S., Famiglietti, J.S., Houser, Trabucco, A., Zomer, R.J., Bossio, D.A., van r a i n b ow, g r e e n, b l u e a n d g r ey
P.R., et al. 2015. The observed state of the Straaten, O. & Verchot, L.V. 2008. Climate water: tree cover and geopolitics of
water cycle in the early twenty-first century. change mitigation through afforestation/ climatic teleconnections. Current Opinion

Unasylva 251, Vol. 70, 2019/1


26

in Environmental Sustainability, 6: 41–47. Wang-Erlandsson, L., Fetzer, I., Keys, P.W., doi.org/10.1002/2016GL072235
https://doi.org/10.1016/j.cosust.2013.10.008 van der Ent, R.J., Savenije, H.H.G. & Weng, W., Costa, L., Lüdeke, M.K.B. &
van Noordwijk, M., Tanika, L. & Lusiana, Gordon, L. J. 2018. Remote land use Zemp, D.C. 2019. Aerial river management
B. 2017. Flood risk reduction and flow impacts on river flows through atmospheric by smart cross-border reforestation. Land
buffer ing as ecosystem ser vices – teleconnections. Hydrology and Earth Use Policy, 84: 105–113. https://doi.
Part 1: Theory on flow persistence, flashiness System Sciences, 22(8): 4311–4328. https:// org/10.1016/j.landusepol.2019.03.010
and base flow. Hydrology and Earth doi.org/10.5194/hess-22-4311-2018 Zemp, D.C., Schleussner, C-F., Barbosa,
Systems Sciences, 21(5): 2321–2340. Wang-Erlandsson, L., van der Ent, R.J., H.M.J., Hirota, M., Montade, V., Sampaio,
https://doi.org/10.5194/hess-21-2321-2017 Gordon, L.J. & Savenije, H.H.G. 2014. G., Staal, A., Wang-Erlandsson, L. &
Wang-Erlandsson, L., Bastiaanssen, Contrasting roles of interception and Rammig, A. 2017. Self-amplified Amazon
W.G. M., Gao, H., Jägermeyr, J., transpiration in the hydrological cycle – Part forest loss due to vegetation-atmosphere
Senay, G. B., van Dijk, A. I. J. M., 1: Temporal characteristics over land. Earth feedbacks. Nature Communications,
G u e r s c h m a n, J. P. , Key s , P.W. , System Dynamics, 5(2): 441–469. https://doi. 8: 14 681. ht t p s://doi. org /10.1038/
Gordon, L.J. & Savenije, H.H.G. 2016. org/10.5194/esd-5-441-2014 ncomms14681. u
Global root zone storage capacity from Wei, Z., Yoshimura, K., Wang, L., Miralles,
satellite-based evaporation. Hydrology D.G., Jasechko, S. & Lee, X. 2017. Revisiting
and Earth System Sciences, 20(4): 1459– the contribution of transpiration to global
1481. https://doi.org/10.5194/hess-20- terrestrial evapotranspiration. Geophysical
1459 2016 Research Letters, 44(6): 2792–2801. https://

Unasylva 251, Vol. 70, 2019/1


27

Dryland forests and agrosilvopastoral


systems: water at the core
A.D. del Campo, M. González-Sanchis, U. Ilstedt, A. Bargués-Tobella
and S. Ferraz
“LEADING GOATS HOME” BY TASECHO IS LICENSED UNDER CC BY-NC-ND 2.0

D
A water-centred approach is ryland systems occur on all con- (PET) in dry subhumid and semiarid lands
essential for maintaining the tinents and cover about 41 per- is considerably higher than annual pre-
resilience of forested drylands in cent of the Earth’s land surface, cipitation, with frequent meteorological
the face of climate change. with little variation in this figure in recent droughts. These atmospheric drivers lead
decades (Cherlet et al., 2018). Drylands to low soil moisture and this, in turn, means
differ in their moisture deficit and can be slow tree growth and low productivity,
Antonio del Campo is Associate Professor
at the Research Institute for Water and classified in four subtypes according to resulting in a socio-ecological context of
Environmental Engineering (Instituto de the United Nations Environment (UNEP) water scarcity. Marked rainfall seasonality,
Ingeniería del Agua y Medio Ambiente – aridity index (AI)1 as dry subhumid (0.65– with torrential events followed by long
IIAMA), Polytechnic University of Valencia,
Spain. 0.5), semiarid (0.5–0.2), arid (0.2–0.05) dry periods, and the combination of high
María González-Sanchis is a researcher at or hyperarid (<0.05) (Figure 1).2 Forests intra- and interannual variability, put such
IIAMA, Polytechnic University of Valencia, and grasslands are the dominant biomes in regions within the “difficult” hydrology
Spain.
Ulrik Ilstedt is Assistant Professor at the the dry subhumid and semiarid subtypes, framework, which hampers water secu-
Department of Forest Ecology and Management, respectively (more than 60 percent of the rity, sustainable development and poverty
Swedish University of Agricultural Sciences. subtype areas). On the other hand, the arid reduction (Grey and Sadoff, 2007).
Aida Bargués-Tobella is a postdoctoral
researcher at the Department of Forest Ecology and hyperarid subtypes are mostly treeless 1
AI is calculated as precipitation (P) divided
and Management, Swedish University of (FAO, 2016) and thus beyond the scope of by PET.
Agricultural Sciences, and at the World this article. 2
Additionally, the Convention on Biological
Agroforestry Centre. Diversity’s delineation includes some areas
Silvio Ferraz is Professor at the Department of Based on their underlying definition (i.e.
with presumed dryland features in which
Forest Sciences, University of São Paulo, Brazil. by AI), annual potential evapotranspiration P/PET > 0.65 (CBD Secretariat, 2010).

Unasylva 251, Vol. 70, 2019/1


28

La Hunde y Palomeras, Spain


!

Saponé, Burkina Faso


!

Jequitinhonha Valley, Brazil


!
Hyperarid
Arid
Semiarid
Dry subhumid
Additional areas included in CBD definition

1
World dryland areas
Note: Dryland categories are as per definitions by the United Nations Convention to Combat Desertification and the Convention on Biological
Diversity (CBD). Black dots show the locations of the case studies.
Source: UNEP-WCMC (2007).

Climate change is expected to cause an southern Africa, Australia, the Middle East Climatic constraints increase the role
increase in the global area of drylands and Central Asia (Cherlet et al., 2018). The of soil processes and properties in the
of 10–23 percent, depending on dryland intensification of precipitation and other regulation and magnitude of water-related
subtype, by the end of the twenty-first climatic extremes under warmer conditions issues in drylands, especially those con-
century, particularly in areas of North is likely to increase water scarcity and cerned with resource storage (e.g. soil
and South America, the Mediterranean, moisture deficits in drylands and beyond. depth, infiltrability, deep-water storage

A combination
of land uses
(e.g. agriculture,
woodlands, pastures
and barren land
shown here) and
management
practices (e.g. soil
treatments and the
© A. DEL CAMPO, 2009

check dam) interact


with climate and soil
processes and affect
the regulation and
magnitude of water-
related issues

Unasylva 251, Vol. 70, 2019/1


29

and erosion). Thus, land-use and manage- index) are important drivers affecting the CASE STUDIES
ment practices, especially nature-based redistribution and subsequent use of water Below, three case studies from drylands
solutions, are extremely important for the in the soil profile. on three continents demonstrate how
soil–water–productivity complex. water-centred management can improve
This article uses case studies in drylands TARGETING WATER IN OBJECTIVES water budgets and local livelihoods,
on three continents to show the importance AND MANAGEMENT OPTIONS increase climate-change resilience
of a water-centred approach to dryland Drylands provide a wide array of goods and adaptation, and reduce the risk
management for increasing resilience and and ecosystem services, but their poten- of disaster.
adaption to climate change. tial is often underestimated because they
are wrongly perceived to be unproductive Pine reforestation in drylands
FOREST ECOSYSTEM PROCESSES (White and Nackoney, 2003). Drylands Monte La Hunde y Palomera (950 metres
AND TREE FUNCTIONAL TRAITS support the livelihoods of more than one- above sea level) is a publicly owned dry-
Dryland forests and agrosilvopastoral third of the global human population by land forest in eastern Spain (Figure 2). The
systems (DFASs) face specific challenges supplying food, forage for livestock and forest covers 4 700 ha, and it includes 887
compared with other vegetation types. Low drinking water. They also provide habitats ha of homogeneous Aleppo pine (Pinus
water availability, low growth and unprece­ for species uniquely adapted to variable halepensis) planted between 1945 and
dented disturbance regimes (e.g. wildfires and extreme environments, which, in turn, 1970 as part of a national afforestation
and pest outbreaks), aggravated by climate constitute sources of genetic material for programme. The Aleppo pine forest has
change, make them less resilient and more developing drought-resistant varieties. high tree density (more than 1 500 trees
prone to shifts toward less-productive Because of their great extent, drylands can per ha, to increase soil protection) and
states (desertification) (Johnstone et al., store large amounts of carbon (Lal, 2004). little silvicultural intervention. The lack of
2016). Anthropogenic pressures (such as The provision of all these goods and intervention is common in many protective
those imposed by grazing, browsing, for- ecosystem services is essentially depen- forests in the Mediterranean.
est overexploitation and deforestation) add dent on water availability, which is often The Monte La Hunde y Palomera forest
complexity and feedbacks. limited, variable and unpredictable but also region has an AI of 0.62, a mean annual
Ecohydrology in drylands is mostly cap- fundamental for supporting flora and fauna. temperature of 13.7 ºC, and precipitation of
tured by the strong relationship between Vegetation dynamics, soil-water flows and 465 mm (1960–2007). The soils are shal-
soil cover and water; that is, forest structure climate are strongly coupled in drylands; low, with high concentrations of carbonate,
– both physical (e.g. tree density, canopy the capacity to cope with temporal water a basic pH, and a sandy-silty loam texture.
cover and basal area) and biological (spe- shortages is essential for both people’s The lack of forest management, com-
cies composition) – has a direct impact livelihoods and the ecosystems them- bined with the climatic characteristics
on water-resource availability (Bosch and selves. Thus, water is the key element for of drylands, has produced a dense forest
Hewlett, 1982), affecting variables such the socio-ecological resilience of drylands in which growth is stalled; it intercepts
as infiltration, evapotranspiration, surface and must constitute a quantitative basis of about 40 percent of gross precipitation and
runoff (and erosion) and groundwater any management approach (Falkenmark, severely competes for the other 60 percent
recharge. On the one hand, decreasing Wang-Erlandsson and Rockström, 2019). (del Campo et al., 2017, and references
canopy cover increases net precipitation, In more humid environments, water yield therein). As a result, the forest is highly
which in turn can increase soil moisture has long been quantified as part of eco- susceptible to climatic fluctuations (i.e.
and related water flows such as ground- system management (Bosch and Hewlett, rainfall variability), thus increasing its
water recharge and water yield, as well as 1982). In dryland ecosystems, water should vulnerability to climate change. Water
soil evaporation. On the other, high tree not just be quantified, it should be central infiltration and percolation is essential not
cover increases interception and transpi- to land planning and management. More only for the forest itself but also for feed-
ration while maximizing soil protection specifically, the emphasis should be on ing two complex aquifer systems, Mancha
and enhancing soil infiltration capacity. soil water and aquifer recharge rather Oriental (7 000 km2) and Alpera (400 km2).
The explicit consideration of trade-offs than on increasing total runoff or stream- The two aquifers comprise the main water
between various hydrological processes flow. Groundwater is the primary water source of 127 000 ha of field crops, but they
and vegetation is essential in drylands resource in drylands because surface water have suffered recurrent drought episodes
when dealing with resource storage (i.e. resources are generally scarce and highly in the last 20 years.
soil and water). Moreover, the water-related unreliable; maximizing its recharge should In this context, the aim of forest manage-
traits of tree species (e.g. canopy and root therefore be targeted as a means to increase ment must be to enhance tree growth and
architecture, wood density, and leaf area the socio-ecological resilience of drylands. vigour (thus reducing the forest’s climatic

Unasylva 251, Vol. 70, 2019/1


30

vulnerability) and soil protection, while


also increasing the catchment water bud-
get and its contribution to downstream
users. Thus, thinning from below at dif-
ferent intensities (higher on flat sites, and
moderate to light on steeper sites) was
performed in a crowded forest, achieving
an alternation of firebreak and groundwater
recharge areas (tree density <170 trees per
ha) together with zones of moderate tree
density (450–700 trees per ha), enough to
promote tree vigour and infiltration without
decreasing soil protection. This manage-
ment approach focuses on soils, trees, water
and climatic factors and can be considered
as ecohydrological-based forest manage-
ment. It has proved capable of coping with
trade-offs among multiple objectives:
© A. CAMARA

canopy interception and stand transpiration


have been reduced; soil water infiltration,
deep percolation, tree transpiration and Aleppo pine forest, Monte La Hunde y
Palomera forest, eastern Spain
water contributions to the aquifers have climate-change-related disturbances.
all increased; and fuel models have been Ecohydrological forest management also sense of insecurity, which is especially
altered. The management changes have has social and economic benefits through- important in the urban–forest interface,
produced a forest with less climatic vul- out the catchment. For example, increasing and potentially avoids the costs of dam-
nerability and lower fire risk (del Campo the water budget increases the capability age caused by wildfire and the expense of
et al., 2017, and references therein) and, of users to cope with drought. Reducing forest restoration. Such benefits arise when
which, therefore, is more capable of facing the fire hazard both decreases the public water is put at the core of the management
approach.
2 Water contribution (mm/yr) Climate dependence (%)
Results of
ecosystem 500 30 Agroforestry parklands:
management,
25 coping with multiple objectives
Monte La Hunde y 400
Palomera 20
but only “one water”
300 Saponé is a rural municipality in central
15
200 Burkina Faso in West Africa. The domi-
10
nant soils are ferric lixisols, with low
nutrient content and sandy-clay and sandy-
100 5
0
U M
0
U M
loam textures. Mean annual precipitation
at Ouagadougou (30 km north of Saponé)
Fire risk (%) Economic revenues (€/ha) was 790 mm in 1952–2014 (in the range of
100 Low 1000 570–1 189 mm). Most rainfall occurs in a
80 800 single rainy season, which runs from April
to October. Mean annual potential evapo-
600
transpiration and mean AI (1974–2003) are
60

40 400 1 900 mm and 0.38, respectively.


20 200 The landscape is characterized by open
High
tree cover (30 trees per ha) dominated by
0 0
U M U M Vitellaria paradoxa (shea), with annual
Notes: Forest water contribution (mm/yr); climate dependence (growth dependence
on previous monthly precipitation in %); fire risk (percentage of days/year with very crops such as pearl millet, sorghum,
high [red], high [orange] and low [green] fire risk); economic revenues (euros per ha),
(see del Campo et al., 2017 for specific references).
groundnut and cowpea grown under and
U = unmanaged, M = managed. among the scattered trees. These cultivated

Unasylva 251, Vol. 70, 2019/1


31

more nuanced story. Soil water drainage


collected at a depth of 1.5 m was high-
est in the area below the edge of the tree
canopy and decreased both towards the
tree stem and towards the centre of adja-
cent open areas among trees (Ilstedt et al.,
2016). Thus, little water was available for
groundwater recharge both close to tree
stems and in the open areas far away from
trees. Interception and transpiration losses
are higher in the area around tree stems,
which explains the reduced deep drainage
in this area. The decrease in water drainage
observed with increasing distance from
the canopy edges of trees towards open
areas, on the other hand, can be attributed
to the observed concurrent decrease in
infiltration capacity and preferential flow
© A. CAMARA

(Bargués-Tobella et al., 2014). Thus, trees


should not be seen only as water consumers
Harvesting aleppo pine, Monte La Hunde y but also as key ecosystem engineers that
Palomera forest, eastern Spain
result of land use, land-cover conversions enable soil and groundwater recharge.
open woodlands are referred to as agro- and human pressure in general; thus, man- In Saponé, groundwater recharge is
forestry parklands, and they constitute agement approaches designed to improve maximized at an intermediate canopy
the predominant farming system in the local livelihoods should aim to increase cover (Ilstedt et al., 2016). At tree cover
Sudano-Sahelian region of West Africa, soil and groundwater recharge. below the optimum, more trees result
covering large areas (Boffa, 1999). Trees Trees consume more water than shorter in increased groundwater recharge
are actively conserved and promoted on vegetation types such as crops and grasses because the improvement in soil hydrau-
farms because of the benefits they provide (Zhang, Dawes and Walker, 2001). Based lic properties conferred by these trees
to local communities – including the provi- on this understanding, increasing tree outweighs additional evapotranspiration
sion of fruits, nuts, shading, medicines, and cover is often discouraged in drylands losses. The opposite is the case, how-
fodder for livestock. because it might jeopardize precious ever, at tree-cover percentages above the
Rainfall is highly variable in Saponé. water resources (Jackson et al., 2005). optimum (Figure 3).
The relatively short rainy season is char- But results from studies conducted in the Although more research is needed, from
acterized by a few intense events unevenly agroforestry parklands of Saponé reveal a a management perspective it is vital to
distributed over time, and there is large
spatial and interannual rainfall variabil-
ity. The soils have low structural stability
and are highly vulnerable to physical
degradation, such as decreases in soil infil-
tration capacity, resulting in limited soil and
groundwater recharge opportunities and a
higher prevalence of infiltration-excess
overland flow. This, in turn, increases the
risk of agricultural drought, erosion and
flooding, placing considerable constraints
on water supply and food production, par-
ticularly given the dominance of rainfed
crops. Physical degradation is typically a
© A. CAMARA

Agroforestry parklands,
Saponé, Burkina Faso

Unasylva 251, Vol. 70, 2019/1


32

Drainage at 1.5 m depth Tree transpiration Groundwater recharge


90 140 60
80
120 50
70
100 40
60

mm/year
mm/year

mm/year
50 80
30
40 60
30 20
40
20
10
10 20

0 0 0
1 10 30 1 10 30 1 10 30
Tree density (trees/ha) Tree density (trees/ha) Tree density (trees/ha)

3
Results of ecosystem management, of rivers draining to the north, northeast, reduces stream flows. In this case, forest
Saponé; three tree densities
southeast and south of the country. The management should be adjusted to water
promote practices that maximize the natural vegetation in the biome has low availability, such as by reducing the area
positive impacts of trees on soil hydraulic biomass density and low interception. This, of plantations, increasing rotation lengths
properties and minimize tree water use and added to the well-drained soils, means (because water use declines with tree age;
interception. Thus, tree species selection, there is a hydric excess responsible for Perry and Jones, 2017), mixing stand ages
tree pruning and livestock control offer recharging aquifers and maintaining (to create a mosaic), and reducing manage-
opportunities to increase groundwater stream flows (Honda and Durigan, 2017). ment intensity.
recharge (Ilstedt et al., 2016). Degradation due to land-use change Another aspect under discussion regard-
(mostly to agriculture) is altering this ing the Cerrado biome, mostly in protected
Cerrado: the hydrological dynamic, however, leading to contaminated areas and remaining fragments, is the
consequences of vegetation streams and reducing water availability. reduction of fire – considered a natural
biomass increment The area of short-rotation forest plan- element of Cerrado ecology (Durigan
The Cerrado is the second-largest biome in tations has increased in the region, with and Ratter, 2016) – brought about by a
Brazil, occupying 204 million ha (24 per- Eucalyptus grandis the most important policy of fire suppression (Durigan and
cent of the country’s total land area); it species. The biomass of these forests Ratter, 2016). Fire reduction is leading to
is subject to considerable land-use pres- increases rapidly, with the trees explor- an increase in vegetation biomass, which
sure (Sano et al., 2019). Vegetation types ing water resources using their deep root in turn results in higher interception and
vary along a regional climatic gradient systems and presenting high evapotrans- modified evapotranspiration dynamics
depending on local soil and geographical piration, potentially altering the soil water (Passos et al., 2018), causing changes in
characteristics and include dry forests, balance. Lima et al. (1990) compared the the hydrological regime and in plant com-
scrub woodland, open scrub (sensu stricto soil water balance in Cerrado vegetation munities. Oliveira et al. (2017) monitored
Cerrado) and grasslands. Annual precipita- with Pinus and Eucalyptus plantations in piezometric wells in various Cerrado
tion is in the range of 1 200–1 800 mm, northeastern Minas Gerais (Jequitinhonha vegetation types over a two-year period
presenting high seasonality (with a six- Valley, annual precipitation = 1 121 mm) and showed that the increase in vegetation
month dry season) and the AI is slightly and showed that the conversion of natural density reduced watertable recharge from
lower than 1. The predominant soils are Cerrado vegetation (36 m3 per ha) to Pinus 363 mm per year (grasslands) to 315 mm
deep, highly weathered and acidic, and they caribaea (210 m3 per ha) and Eucalyptus per year (Cerrado). Differences in evapo-
have low nutrient concentrations. Because grandis (366 m3 per ha) increased intercep- transpiration rates and soil water content
nutrient deficiency can be corrected, and tion losses by 74 mm per year for Pinus and were also observed among vegetation types
other soil characteristics are highly favour- 134 mm per year for Eucalyptus (Figure 4). (Miranda et al., 2003).
able, some lands in the Cerrado have been The soil water balance decreased from Land-use change in the Cerrado, and in
converted to agriculture; production is high 556 mm per year in natural Cerrado veg- other drylands worldwide, requires taking
when fertilizers are used. The Cerrado, etation to 450 mm in Pinus and 326 mm into account the hydrological constraints
therefore, has become one of the world’s in Eucalyptus. The reduction in water (made clear by the characteristics of the
most threatened biodiversity hotspots availability by plantations increases the natural vegetation) to maintain hydro-
(Klink and Machado, 2005). effects of natural seasonality (i.e. lower logical processes and the provision of
The Cerrado concentrates the headwaters availability during dry seasons) and ecosystem services.

Unasylva 251, Vol. 70, 2019/1


33

Water contribution (mm/yr) ET (% of Gr) Interception (% of Gr)

90 14
600
80
12
500 70
10
60
400
50 8
300
40 6

200 30
4
20
100 2
10

0 0 0
P E C P E C P E C
4
Comparison of plantations and natural Cerrado vegetation, Jequitinhonha Valley, Brazil
Note: Water contribution (mm/yr); evapotranspiration (% of gross precipitation – Gr); interception (% of Gr). C = Cerrado, E = Eucalyptus, P = Pinus.

CHALLENGES IN THE can contribute to several Sustainable in dryland forests and agrosilvopastoral
GOVERNANCE AND MANAGEMENT Development Goals (SDGs), including systems are not clearly marketable, discour-
OF DRYLAND FORESTS AND SDG 2 (Zero hunger), SDG 6 (Clean aging potential investment in management.
AGROSILVOPASTORAL SYSTEMS water and sanitation) and SDG 15 (Life on Decision-support systems capable of
Water plays a fundamental role in socio- land). But it is a highly complex challenge, handling complexity and multiple inter-
ecological resilience (Falkenmark, with economic, social, environmental actions, and which might encompass
Wang-Erlandsson and Rockström, 2019), and climatic dimensions. The need for economic valuation (Tecle, Shrestha
particularly in drylands. Forward-looking multiple goods and services increases and Duckstein, 1998), present a
governance and management policies in the complexity of the challenge because potential means for negotiating the com-
dryland forests and agrosilvopastoral sys- their quantity, typology and valuation plexity of water-oriented land management
tems, therefore, need to consider water (in economic terms) vary with ecosys- in drylands.
as a crucial supporting element for the tem type (La Notte et al., 2015) and
production of goods and services, at least hamper the potential for a generalized CONCLUSION
at the same level as biomass and carbon. approach applicable to all dryland for- The water-oriented management of dryland
Water-oriented land management ests and agrosilvopastoral systems. forests and agrosilvopastoral systems may
Also, many of the products produced increase water availability and therefore
The Cerrado, Brazil socio-ecological resilience. As the case
studies presented above show, strate-
gies such as canopy opening, pruning
and species selection can be effective in
combating water scarcity (by increas-
ing soil and groundwater recharge)
while also increasing climate-change
resilience and adaptation. The optimum
management intensities and strategies are
likely to vary with ecosystem character-
istics, even within the same catchment
or region.
The need to provide multiple goods and eco-
system services increases the management
challenge but also the potential benefits
and therefore management possibilities.
The complexity of multi-objective
management approaches, and the eco-
© IIAMA

logical variability of dryland forests

Unasylva 251, Vol. 70, 2019/1


34

CERRADO (CAVERNA AROE JARI) BY STINKENROBOTER IS LICENSED UNDER CC BY-NC 2.0


A patch of Cerrado surrounded by farm fields
on the access road to Caverna Aroe Jari, on preferential flow and soil infiltrability Bautista, I., Ruiz, G. & Francés, F. 2017.
Mato Grosso, Brazil in an agroforestry parkland in semiarid Ecohydrological-based forest management
Burkina Faso. Water Resources Research, in semi-arid climate. In: J. Křeček, M.
and agrosilvopastoral systems, means 50(4): 3342–3354. Haigh, T. Hofer, E. Kubin & C. Promper, eds.
that more effort is needed to quantify Boffa, J.M. 1999. Agroforestry parklands in E cos ystem ser vices of h ea d wa ter
and value the goods and ecosystem sub-Saharan Africa. FAO Conservation catchments, Chapter 6. Springer International
services of dryland forests and agrosilvo- Guide 34. Rome, FAO. Publishing and Capital Publishing Co.
pastoral systems and to incorporate this Bosch, J.M. & Hewlett, J.D. 1982. A review Durigan, G. & Ratter, J.A. 2016. The
information in management. of catchment experiments to determine need for a consistent fire policy for
the effect of vegetation changes on water Cerrado conservation. Journal of Applied
ACKNOWLEDGEMENT yield and evapotranspiration. Journal of Ecology, 53(1): 11–15.
The authors thank CEHYRFO-MED Hydrology, 55(1–4): 3–23. Falkenmark, M., Wang-Erlandsson, L. &
(CGL2017-86839-C3-2-R), LIFE17 CCA/ CBD Secretariat. 2010. Decision adopted Rockström, J. 2019. Understanding of water
ES/000063 RESILIENTFORESTS, by the Conference of the Parties to the resilience in the Anthropocene. Journal of
Swedish Research Council Formas (2017- Convention on Biological Diversity [CBD] Hydrology, X, 2: 100009.
00430), and Swedish Research Council VR at its tenth meeting: X/35 Biodiversity of FAO. 2016. Trees, forests and land use in
grant 2017-05566. u dry and sub-humid lands. Conference of drylands. The first global assessment.
the Parties to the Convention on Biological Preliminary findings. Rome.
Diversity, Tenth meeting. Nagoya, Japan, Grey, D. & Sadoff, C.W. 2007. Sink or swim?
18–29 October. Water security for growth and development.
Cherlet, M., Hutchinson, C., Reynolds, J., Water Policy, 9: 545–571.
Hill, J., Sommer, S. & von Maltitz, G., Honda, E.A. & Durigan, G. 2017. A
References eds. 2018. World atlas of desertification. restauração de ecossistemas e a produção
Luxembourg, Publication Office of the de água. Hoehnea, 44(3): 315–27.
Bargués-Tobella, A., Reese, H., Almaw, European Union. Ilstedt, U., Bargués-Tobella, A., Bazié,
A., Bayala, J., Malmer, A., Laudon, H. del Campo, A.D., González-Sanchis, M., H.R., Bayala, J., Verbeeten, E., Nyberg,
& Ilstedt, U. 2014. The effect of trees Lidón, A., García Prats, A., Lull, C., G., et al. 2016. Intermediate tree cover can

Unasylva 251, Vol. 70, 2019/1


35

maximize groundwater recharge in the and natural “Cerrado” vegetation measure Coutinho, A., et al. 2019. Land use dynamics
seasonally dry tropics. Scientific Reports, by the soil water balance method. IPEF in the Brazilian Cerrado in the period from
6: 21930. International, 1: 35–44. 2002 to 2013. Pesquisa Agropecuária
Jackson, R.B., Jobbágy, E.G., Avissar, R., Miranda, A.C., Lloyd, J., Santos, A.J.B., Brasileira, 54(0).
Roy, S.B., Barrett, D.J., Cook, C.W., Silva, G.T.D.A. & Miranda, H.S. 2003. Tecle, A., Shrestha, B.P. & Duckstein, L.
Farley, K.A., le Maitre, D.C., McCarl, B.A. Effects of fire on surface carbon, energy and 1998. A multiobjective decision support
& Murray, B.C. 2005. Trading water for water vapour fluxes over campo sujo savanna system for multiresource forest management.
carbon with biological carbon sequestration. in central Brazil. Functional Ecology, 17(6): Group Decision and Negotiation, 7(1):
Science, 310(5756): 1944–1947. 711–719. 23–40.
Johnstone, J.F., Allen, C.D., Franklin, J.F., Oliveira, P.T.S., Boccia Leite, M., Mattos, UNEP-WCMC. 2007. A spatial analysis
Frelich, L.E., Harvey, B.J., Higuera, P.E., T., Nearing, M.A., Scott, R.L., de Oliveira approach to the global delineation of dryland
et al. 2016. Changing disturbance regimes, Xavier, R., da Silva Matos, D.M. & areas of relevance to the CBD Programme of
ecological memory, and forest resilience. Wendland, E. 2017. Groundwater recharge Work on Dry and Subhumid Lands. Dataset
Frontiers in Ecology and the Environment, decrease with increased vegetation density in based on spatial analysis between WWF
14(7): 369–378. the Brazilian Cerrado. Ecohydrology, 10(1). terrestrial ecoregions and aridity zones.
Klink, C.A. & Machado, R.B. 2005. https://doi.org/10.1002/eco.1759 Dataset checked and refined to remove many
Conservation of the Brazilian Cerrado. Passos, F.B., Schwantes Marimon, B., gaps, overlaps and slivers (July 2014).
Conservation Biology, 19(3): 707–713. Phillips, O.L., Morandi, P.S., Carvalho White, R.P. & Nackoney, J. 2003. Drylands,
La Notte, A., Liquete, C., Grizzetti, B., das Neves, E., Elias, F., Reis, S.M., de people, and ecosystem goods and services: a
Maes, J., Egoh, B. & Paracchini, M. 2015. Oliveira, B., Feldpausch, T.R. & Marimon web-based geospatial analysis. Washington,
An ecological-economic approach to the Júnior, B.H. 2018. Savanna turning into DC, World Resources Institute.
valuation of ecosystem services to support forest: concerted vegetation change at the Zhang, L., Dawes, W. & Walker, G. 2001.
biodiversity policy: a case study for nitrogen ecotone between the Amazon and ‘Cerrado’ Response of mean annual evapotranspiration
retention by Mediterranean rivers and lakes. biomes. Revista Brasileira de Botanica, 41(3) to vegetation changes at catchment
Ecological Indicators, 48: 292–302. https://doi.org/10.1007/s40415-018-0470-z scale. Water Resources Research, 37(3):
Lal, R. 2004. Carbon sequestration in dryland Perry, T.D., & Jones, J.A. 2017. Summer 701–708. u
ecosystems. Environmental Management, streamflow deficits from regenerating
33(4): 528–544. Douglas-fir forest in the Pacific Northwest,
Lima, W.P., Zakia, M.J.B., Libardi, P.L. USA. Ecohydrology, 10(2): e1790.
& Souza Filho, A.P. 1990. Comparative Sano, E.E, Rosa, R., de Mattos Scaramuzza,
evapotranspiration of Eucalyptus, pine C.A., Adami, M., Bolfe, E.L., Camargo

Unasylva 251, Vol. 70, 2019/1


36

Gaps in science, policy and practice


in the forest–water nexus
M. Gustafsson, I. Creed, J. Dalton, T. Gartner, N. Matthews, J. Reed,
L. Samuelson, E. Springgay and A. Tengberg

More knowledge is needed


urgently on how to implement
effective management regimes for
the forest–water nexus.

Malin Gustafsson is Programme Officer at the


Swedish Water House, Stockholm International
Water Institute (SIWI), Stockholm, Sweden.
Irena Creed is Professor at the School of
Environment and Sustainability, University of
Saskatchewan, Saskatoon, Canada.
James Dalton is Director of the Global
Water Programme, International Union for
Conservation of Nature (IUCN), Gland,
Switzerland.
Todd Gartner is Director of Cities4Forests
and the Natural Infrastructure Initiative at the
World Resources Institute, Washington, DC,
United States of America.
Nathanial Matthews is Program Director at
the Global Resilience Partnership hosted at
the Stockholm Resilience Centre, Stockholm
University, Stockholm, Sweden.
James Reed is Researcher in the Sustainable
Landscapes and Food Systems team at the
Center for International Forestry Research,
Bogor, Indonesia, and at the University of
Cambridge Conservation Research Institute,
Cambridge, United Kingdom of Great Britain
and Northern Ireland.
Lotta Samuelson is Programme Manager at
the Swedish Water House, SIWI, Stockholm,
Sweden.
Elaine Springgay is Forestry Officer at the
FAO Forestry Department, Rome, Italy.
Anna Tengberg is Program Manager at the
Swedish Water House, SIWI, Stockholm,
Sweden, and Adjunct Professor at Lund
University Centre for Sustainability Studies,
"TREE SCIENCE-9438" BY SOME.DUDE IS LICENSED UNDER CC BY-NC-SA 2.0

Lund University, Lund, Sweden.


 

The crucial role of


trees and forests in
hydrologic cycles
requires more
research

Unasylva 251, Vol. 70, 2019/1


37

T
rees and forests play important INCREASING ATTENTION ON sectors – including those between forests
roles in hydrologic cycles, such INTERLINKAGES and water – can be realized (Seddon et
as by altering the release of water The need for greater integration among al., 2019).
into the atmosphere, influencing soil mois- sectors and scales in natural resource gov- The productivity of multifunctional land-
ture and improving soil infiltration and ernance is now widely recognized (e.g. scapes is contingent on the management of
groundwater recharge (Springgay et al., Liu et al., 2018), such as by incorporating interlinked forest and water processes and
2018). Forest-related changes in land use stakeholder-driven, bottom-up approaches resources (Ilstedt et al., 2007), which, in
such as deforestation, reforestation and into natural resource management strat- turn, requires adequate knowledge of such
afforestation can affect both nearby and egies (Creed and van Noordwijk, 2018; linkages. But persistent knowledge gaps
distant water supplies (Jones et al., 2019): Tengberg and Valencia, 2018). Landscape on forest–water interactions require atten-
for example, a decrease in evapotranspira- approaches are being discussed and, in tion and action if system-based approaches
tion following deforestation in one area some cases, adopted in both tropical and such as forest and landscape restoration
may reduce rainfall in downwind areas temperate regions in an attempt to rec- (Laestadius et al., 2015; Carmenta and
(Ellison et al., 2017). Climate change and oncile often conflicting environmental Vira, 2018) and landscape approaches
an increase in extreme weather events are and developmental challenges at broader that holistically consider the importance
disturbing water cycles and threatening spatial scales (Estrada-Carmona et al., of healthy forests for sustainable water
the stability of water flows (IPCC, 2019). 2014; García-Martín et al., 2016; Reed supplies, and vice versa, are to be applied.
Meanwhile, water supplies are affected by et al., 2016). Managing the forest–water nexus is
an increase in human water consumption to Forests and water – both of which are integral to achieving many of the SDGs
meet domestic, agricultural and industrial integral landscape components – are also and could be better acknowledged in the
needs (Rockström et al., 2009). Increasing receiving increased attention in policy implementation of (intended) nation-
demand for water is reducing freshwater and practice. The European Union Water ally determined contributions under the
flows and groundwater levels, often with Framework Directive, adopted in 2000, has Paris Agreement on climate change. The
negative effects on biodiversity, ecosystems been a strong driver of bottom-up public SDGs provide a useful framework for
and ecosystem services (Power, 2010). and private partnerships to secure water bringing this nexus to the attention of
Water, forests and climate, therefore, are quality and flows. Globally, the Convention policy- and decision-makers. We suggest
intrinsically interlinked at multiple levels on Wetlands of International Importance, that, to effectively implement large-scale
in what has been termed the forest–water especially as Waterfowl Habitat (generally landscape approaches, adequate attention
nexus, but the underlying systems and feed- known as the Ramsar Convention) has must be given to the multifunctionality
back loops in forest hydrologic processes raised awareness of the need to conserve of landscapes, including the interactions
are poorly understood – and also poorly and sustainably use wetlands, forests and and interdependencies of actions across
represented in policy- and decision-making water resources (Tengberg et al., 2018). typically conflicting sectors and their
(Creed and van Noordwijk, 2018). At the The SDGs address water in SDG 6 (Clean importance for livelihoods, climate and
national level, pervasive policy challenges water and sanitation) and forests in SDG 15 economic resilience.
are often confounded by a lack of collabo- (Life on land), although it has been argued
ration among key sectors, such as forestry, that all the SDGs are relevant to forest FOREST–WATER INTERLINKAGES:
water, energy and agriculture. Complex and water management and use (Creed AREAS OF AGREEMENT
multisectoral issues such as forest–water and van Noordwijk, 2018). An increasing AND DISAGREEMENT, AND
relations, therefore, tend to be overlooked focus on forests is evident in other recent KNOWLEDGE GAPS
and inadequately incorporated into sectoral international conventions and campaigns, Forests and water are linked through their
policies. such as the Bonn Challenge, which aims multiple functions, such as the regulation
In this article, we draw on our collec- to restore 350 million ha by 2030, the of basin flows, the reduction of floods and
tive experience and the recent literature to New York Declaration on Forests, which droughts, and the impacts of forests on
highlight knowledge gaps in the integration has the goal of ending deforestation by water yield and quality. There is limited
of the forest–water nexus in science, policy 2030, and the Trillion Trees Partnership, knowledge, however, on the factors that
and practice, including the climate-change which aims to protect and restore 1 trillion regulate these multiple functions, their
discourse and the Sustainable Development trees by 2050. Although the momentum interactions, and ultimately their effects
Goals (SDGs). for and pledges towards these endeavours on those relying on them for water and
have been significant, challenges remain other ecosystem services. The complexity
in translating them into action, and it is of highly contextualized forest–water rela-
still unclear whether synergies across tionships requires management decisions

Unasylva 251, Vol. 70, 2019/1


38
“TROPICAL HORSESHOE BEND” BY THE_OLIVERA IS LICENSED UNDER CC BY-NC-SA 2.0

based on science and an understanding of assessments and the design of supportive as web-based multistakeholder meetings
cross-scalar local–national–global con- policies, to dialogue and communication and information sharing via social media
ditions (Eriksson et al., 2018). Figure 1 (Table 2). Challenges remain, however, in and public fora. Maps explaining water
and Table 1 identify topics where there is overcoming siloed or sectoral approaches yields by ecosystem type, forest cover
both agreement and disagreement among to enable the implementation of integrated and other parameters would be useful
experts, suggesting a need for further forest and water management. There is in understanding and explaining forest
investigation and discussion. For example, often poor communication and a lack of hydrology as part of water management,
there is consensus that the hydrologic trust among stakeholders, as well as a lack together with decision-support tools link-
processes that are influenced by forests of economic incentives for the sustainable ing science, policy and practice.
can influence the water cycle, but there is management of the forest–water nexus.
disagreement on the impacts of these inter- Therefore, we encourage new ways of FOREST, WATER, CLIMATE, AND A
actions. Ongoing research spans a wide communicating the results of scientific LANDSCAPE PERSPECTIVE
range of topics, from technical aspects research using a common terminology and Climate change has many fundamental
related to (for example) water budgets, to modern communication approaches, such and increasing impacts on the global water

Unasylva 251, Vol. 70, 2019/1


39

1
Areas for further discussion within the
forest–water nexus

RELATIONSHIPS BETWEEN
FORESTS, PRECIPITATION WATER VAPOUR TRANSPORT
AND FLOOD CONTROL
PRECIPITATION

EXTENT
EXTENT OF
OF THE
THE EFFECTS
EFFECTS OR
OR
IMPLICATIONS
IMPLICATIONS OF
OF FOREST
FOREST OPTIMAL LOCATION FOR
EVAPOTRANSPIRATION
EVAPOTRANSPIRATION AND
AND AND SPACING OF TREES
PRECIPITATION
PRECIPITATION RECYCLING
RECYCLING IN THE LANDSCAPE

EVAPOTRANSPIRATION

EVAPORATION
TYPES OF
TREES/FORESTS THAT
IMPACTS OF MOST EFFICIENTLY
REFORESTATION SUPPORT WATER
ON WATER YIELD SUPPLY/ SECURITY OR PARAMETERS RELATED TO
DEPLETE WATER YIELD A FOREST'S POTENTIAL
TO MITIGATE FLOODS
INFILTRATION

STREAM FLOW

SURFACE RUNOFF

GROUNDWATER FLOW
BENEFITS/TRADE-OFFS OF FOREST– WATER RESILIENT AND PRODUCTIVE
INTERACTIONS ON ECOSYSTEM SERVICES, LANDSCAPES THAT MAXIMIZE THE
SUCH AS CARBON SEQUESTRATION AND THE POSITIVE AND NEGATIVE BENEFITS OF ECOSYSTEM SERVICES
STORAGE, BIODIVERSITY CONSERVATION, IMPACTS OF FORESTS ON FROM THE FOREST–WATER NEXUS
NUTRIENT CYCLING, SEDIMENTS, DOWNSTREAM WATERFLOW FOR ENVIRONMENTAL AND
AND SOIL-WATER RETENTION AND GROUNDWATER LIVELIHOOD SECURITY

Note: This conceptual overview and summary of issues in need of further discussion within the forest–water nexus arose from discussions at the first Forest Water Champions
workshop in Stockholm, August 2017.

cycle and regional weather patterns (Hegerl on climate change (Seddon, 2018), and management of the forest–water nexus.
et al., 2015), yet water is rarely visible in ecosystems are increasingly managed for To ensure resilient and productive multi-
negotiations within the United Nations water security. But lessons arising from functional landscapes, there is a need to:
Framework Convention on Climate Change NBS approaches, such as projects on • improve understanding of hydrologic
(UNFCCC). Opportunities have been ecosystem-based adaptation, have been processes at a landscape scale;
missed in managing the forest–water nexus inconclusive to date due to monitoring • support the development of new
for climate-change adaptation and mitiga- difficulties and a lack of interventions at integrated knowledge that can
tion, for example by including nature-based a meaningful scale beyond community underpin evidence-based decision-
solutions (NBSs) (Tengberg et al., 2018). projects (Reid et al., 2019). making related to landscapes, forests
At the national level, developing coun- The understanding and consideration and water;
tries have taken the lead in incorporating of water and hydrologic processes can be • strengthen multilevel governance
NBSs in their nationally determined con- used as a key entry point in promoting arrangements that enable genuine
tributions (NDCs) to the Paris Agreement landscape restoration and the sustainable stakeholder participation;

Unasylva 251, Vol. 70, 2019/1


40

TABLE 1.
Areas of agreement, and areas for further discussion, identified at the first Forest Water Champions workshop in
Stockholm, August 2017

Theme Areas of agreement Areas for further discussion

Water quantity Trees and forests influence the hydrologic cycle by regulating and affecting basin • Positive and negative impacts of forests on
flows through interception, uptake, evapotranspiration, reducing runoff and downstream waterflow and groundwater levels
improving soil infiltration and groundwater recharge
The effect of forests on water yield (positive or negative) is dependent on • Types of trees/forests that most efficiently support
location, forest type and age, and scale (physical and temporal) water supply/security
• Types of trees/forests that deplete water yield
• Optimal locations for trees in landscapes
Fire is a normal and healthy aspect of many forests, correlating with
precipitation regimes and influencing hydrology
Forests can reduce the risk of flooding, but the parameters for reduced flood • Parameters related to the potential of forests to
risk are complex, influenced by many factors, and not well known mitigate floods

Evapotranspiration from forests can have a positive effect on downwind • Extent of the effect and implications of forests,
precipitation evapotranspiration and precipitation recycling
There is a need to define the parameters of forest–water relationships • Relationships between forests, precipitation and
flood control
• Impacts of reforestation on water yield
• Benefits and impacts of forests at different scales
Water quality Forests generally improve water quality through their root systems and
stable soil profiles, which can act as natural filters, reducing soil erosion and
sedimentation

Forests are crucial for aquatic ecosystems • Impacts of riparian zones and floodplains on fishing
communities
Policy and The SDGs provide an opportunity to:
practice • bring related scientific knowledge to the attention of policy- and
decision-makers
• highlight areas where the science is not sufficiently comprehensive or is
overly simplistic
• highlight the need for integrated approaches across sectors and disciplines

• Forests are adapted to environmental conditions – including water • Benefits and trade-offs related to forests and water
• Forest–water relationships are scale- and context-dependent interactions – such as carbon storage, climate-
• The impacts of forest and land management activities on water and water change mitigation and adaptation, and extreme
users depend on local conditions, forest ecology, management regime, scale, events
etc. • Development and communication of research,
methods and decision-making tools
The following are needed to improve the management of the forest–water • How to achieve these and by whom
interface:
• a combination of technical and policy measures, summarized in a theory of
change
• a scientific conceptual framework with the main linkages and interactions
between forests and water
• the integration into existing tools to manage for uncertainty/risk in
sustainable forest management, sustainable land management and
integrated water resource management
• communication across disciplines and sectors to enable consensus
Socio-economics Forest and water resources are part of deeply intertwined socio-ecological • How to best approach and manage a mosaic of
systems. Thus, the socio-economic dimensions and implications for governance land uses and other interventions, including natural
policies need to be better addressed, with specific attention to climate change, ecosystems and managed systems, to maximize
reduced forest functions and increased demand for water for human well-being overall benefits, keeping in mind the equitable
distribution of benefits

Ecosystem services from the forest–water nexus need to be better documented, • Whether water accounting needs to re-prioritize
accredited and used to develop funding schemes for overall landscape ecosystems and look at a water “net balance”. This
development can contribute to a cost–benefit analysis of forests
as natural capital in place of or to complement grey
infrastructure

Note: The first Forest Water Champions workshop was organized by FAO, IUCN and SIWI.
Source: Adapted from Springgay et al. (2018).

Unasylva 251, Vol. 70, 2019/1


41

TABLE 2.
Summary of group discussions at a parallel session during the IUFRO Joint Conference on Forests and Water held in
Valdivia, Chile, in November 2018

Ongoing relevant Vulnerability assessments related to forest-watershed management that integrate science, policy and stakeholder engagement
research
Research on payments for ecosystem services and how to design supportive policies using social network analysis
Support for communication, the compilation of indicators and stakeholder initiatives
Water budgets in forestry and agriculture to increase irrigation efficiency
Soil management using native forests and subsidies for some activities
Development of strategies to work with local communities and facilitate discussions between companies and communities
Remaining There is a siloed approach to water – people need to be brought together, but this requires resources, engagement and
challenges leadership as well as a platform
Building trust among different stakeholders is key. There is, for example, a lack of transparency in companies and a reluctance
to share data and information. Landowners do not trust scientists, whom they believe are on the “government’s side”. Non-
governmental organizations have an important role to play in building trust between practitioners, policymakers and scientists
Scientists, policymakers and practitioners speak different “languages”. Scientists and researchers always sound unsure. Why
do they keep saying, “We know this but we don’t know that”?
It is important that there is a common objective; this should be clear from the start but is often missing. What are we trying to
achieve, and what problems are we trying to solve?
Economic incentives for the sustainable management of the forest–water nexus are lacking
Communication New ways of communicating results must be developed. For example, cartoons could be used to increase understanding of
needs the forest–water nexus by using more pictures and less text. Webinars to raise awareness could be organized
There is a need for a common terminology; this could be included in the common objective
It is important to communicate the benefits of addressing the forest–water nexus. Why is it needed, and what will we gain?
Decision-support tools could be developed jointly by scientists, practitioners and policymakers
Note: The parallel session was co-organized by FAO, SIWI and the Swedish Forest Agency.

• identify and apply best management applied – to performance, equity and to restore degraded forests and landscapes
practices and innovative tools for the power dynamics (Kusters et al., 2018; at scale seem remote. Water stress in many
sustainable management of forests and Morrison et al., 2019). The establishment countries has been linked to land degra-
water in landscapes; and or refinement of existing multistakeholder dation resulting from forest conversion
• ensure adequate long-term financing fora can serve to better communicate the (Curtis et al., 2018; IPCC, 2019), and the
for landscape approaches that sustain co-benefits of sustainably managing the growing competition for land between,
ecosystem services and support liveli- forest–water nexus for climate, landscapes for example, agriculture, industries and
hoods (Tengberg et al., 2018). and people; increase awareness of these intensive forestry adds to the challenge.
Achieving successful and inclusive benefits among policymakers, civil society Below, we outline the key implications and
multistakeholder dialogue will inevita- and the general public; and inspire the opportunities for science, policy and prac-
bly require a reconsideration of existing co-production of integrated forest–water tice in addressing the forest–water nexus.
governance structures and institutional knowledge and solutions.
interplays and a move towards more Science
cross-cutting multilevel1 or polycentric2 IMPLICATIONS FOR SCIENCE, There is consensus around many of the
structures (e.g. Ostrom, 2010; Nagendra POLICY AND PRACTICE physical processes that change the hydro-
and Ostrom, 2012). Although such struc- The world continues to lose primary logic cycle and are influenced by forests;
tures have been widely advocated and forests: 2017 has been described as the for example, forests affect water infiltration
supported in recent years, sufficient atten- second-worst year for tropical tree loss and soil hydraulic properties (Neary, Ice
tion is required – and relevant frameworks on record (Curtis et al., 2018), and cur- and Jackson, 2009), and evapotranspira-
rent trends and forecasts in the Brazilian tion from forests can influence downwind
1
i.e. Reconciling governing bodies across Amazon are of major concern (INPE, precipitation (Ellison et al., 2017). Existing
horizontal or vertical scales of influence, undated). The continuation of deforesta- data and knowledge can assist in priori-
from local to national.
tion directly conflicts with the ambitions tizing landscape management strategies
2
i.e. Reconciling governing bodies across
spatial scales, with multiple centres of and commitments embodied in the SDGs and in identifying water-related eco­system
decision-making in a landscape. and other global goals and makes the quest services such as soil erosion control, flood

Unasylva 251, Vol. 70, 2019/1


42

© ANNA TENGBERG/SIWI
A protected forest
in Serra do Mar,
southeastern Brazil,
which helps ensure
a sustainable water
supply to São Paulo

reduction and groundwater recharge. require additional information on water societies (Reed et al., 2017). The impor-
The same data can be useful in identi- resources, including current water avail- tance of these ecosystem services, and
fying trade-offs where the establishment ability and predictions of future changes of sustainably managing the relationship
of forests may be counterproductive to in availability due to climate change and between forests and water, needs to be
water needs. Hydrologic processes in for- human and economic development needs. better recognized in forestry, forest and
ests change over time (Filoso et al., 2017), An enhanced understanding of current water management strategies, and climate
and trees have long lifecycles. Long-term and future water availability and flows and landscape restoration policies, invest-
planning is essential, therefore, for forest would help improve the contributions of ments and initiatives. Encouragingly, many
and landscape restoration, and this can be forest management and restoration inter- recent international sustainability agendas
aided by mapping the potential impacts ventions to water resources (Eriksson et (e.g. those of the UNFCCC, the SDGs and
of climate-change projections on water. al., 2018). For example, the effectiveness the Convention on Biological Diversity)
An important task is to invest in stud- of restoration in storing carbon is partly explicitly call for the increased integration
ies to identify the range of forest–water dependent on adequate soil moisture. of sectors, and the policy environment in
interactions and determine how processes International policy processes focused many countries appears to be changing
and effects occur at different spatial and on halting deforestation, preventing forest to embrace the concept of integration
temporal scales to enable the drawing of degradation and restoring forests are also and to call for engagement across min-
general conclusions on suitable manage- important for – as well as reliant on – water istries (O’Connor et al., forthcoming).
ment approaches. Local parameters such security. Interventions must therefore be Nevertheless, a more specific focus on the
as geography, altitude, forest type, man- context sensitive, and they should take into forest–water nexus is required in national
agement regime, scale and season need account technical data on tree species’ and subnational policy development. We
to be considered in efforts to improve traits and adaptations (Ilstedt et al., 2016). consider that there is sufficient reliable
understanding of highly contextualized information on basic forest–water inter-
forest–water relationships (Creed and Policy linkages to start aligning policy around
van Noordwijk, 2018). For example, cur- Forest–water provisioning, regulating and them – although awareness is needed of
rent investments in landscape restoration cultural ecosystem services are crucial for knowledge gaps, and support processes

Unasylva 251, Vol. 70, 2019/1


43

will be important for improving knowledge and practice to understand and sustainably and communities within broader
and related policies. manage the forest–water nexus. We rec- socio-ecological systems.
New institutional and governance frame- ognize, however, that important gaps exist • In planning future activities, take into
works can play key roles in optimizing in current understanding and application. account likely changes in hydrology
forest–water management in the face of To our knowledge, no method exists for due to climate change.
changing climatic conditions, and a cross- monitoring how changes in landscapes,
sectoral approach is fundamental. The including forest losses and gains, relate Non-governmental and
security of water, energy and food should to changes in water (and vice versa) at the intergovernmental organizations
be core components of forest management scale of a river basin and higher (e.g. in • Analyse the extent to which forest–
and the restoration of multifunctional moisture transfer and precipitation form- water interlinkages are recognized in
landscapes aimed at achieving SDG 15. ing). A lack of data reduces the capacity NDCs, commitments on forest and
Water management would benefit from of managers and policymakers to make landscape restoration, the Aichi Bio-
better integration with forest management informed, evidence-based decisions. There diversity Targets and the post-2020
as an NBS to help achieve SDG 6. These is an urgent need to design, implement and Biodiversity Framework.
aspects should be incorporated into NDCs learn from landscape approaches that both • Initiate dialogues with donors on
and national-level action plans for imple- rely on and influence relationships between financing the implementation of major
menting the 2030 Agenda for Sustainable forests and water. We consider this to be forest and water management activi-
Development. fundamental for accelerating progress ties under the SDGs, such as restora-
towards sustainability goals. Below, we tion initiatives and the management
Practice make specific recommendations for future of water ecosystems.
Progress in sustainable development is research and action.
highly dependent on the management of Communication – all stakeholders
the forest–water nexus, given its impli- Decision-makers • Communicate results using a common
cations for carbon storage, livelihoods • Integrate forest–water interlinkages terminology and modern information
and biodiversity; thus, it is necessary into the climate-change discourse, and communication technologies to
to manage forests for water as well as where the concept of “water resilience” reach a wider range of stakeholder
for biodiversity, climate and economic could help bridge multiple sectoral groups and sectors.
development. In many cases, this will interests (e.g. forests, water, energy and • Reach out to stakeholders at all levels
require a fundamental revision of forest agriculture) (Rockström et al., 2014). and connect local and national-level
management practices, as well as strong • Give greater attention to the role of stakeholders to relevant international
communication, awareness raising and water in climate-change mitigation and fora and dialogue processes relevant
capacity building. There is also a need to adaptation in UNFCCC negotiations, to the forest–water nexus, such as the
develop new institutional frameworks that the Paris Agreement Road Map, and multilateral environmental agreements
foster collaboration on forest and water the NDCs. and World Water Week.
management, such as source-to-sea insti- • Promote science-based management
tutions that address entire management with an understanding of spatial scale ACKNOWLEDGEMENTS
chains (Liss Lymer, Weinberg and Clausen, at multiple levels (i.e. local–national– This article is the result of the Forest Water
2018). Governance considerations, includ- global conditions), as well as issues Champions (FWC) network meetings
ing justice, equity, gender, and indigenous of performance, equity and power co-organized by FAO, IUCN and SIWI.
rights and knowledge are also crucial for dynamics. The first meeting, in August 2017, hosted
managing the forest–water nexus. We • Promote the quantified assessment of 12 experts from the forestry and water
propose the development of environmen- performance of integrated landscape sectors from Europe and North America.
tal and socio-economic multiple-benefits approaches to study activities and The group of experts had the ambition
case studies to help unpack the complexity processes over time, including forest– to establish a consensus on forest–water
of the forest–water nexus and to clarify water functions and interactions and issues and to identify joint future activities
its multiple benefits for the provision of their socio-ecological effects. to promote the forest–water nexus. The
ecosystem services. meeting led to the development of a joint
Restoration managers/practitioners statement on forests and water, as well as a
CONCLUSION AND NEXT STEPS • Base future activities on landscape report of suggested collaborative activities
Attention is increasingly being paid to approaches that recognize the inter- (Springgay et al., 2018). A second FWC
integrated approaches in science, policy linkages of land uses, natural resources meeting was held the following year, in

Unasylva 251, Vol. 70, 2019/1


44

which some of the original FWC experts Curtis, P.G., Slay, C.M., Harris, N.L., analysis. Forest Ecology and Management,
were joined by others. FWC 2 was offi- Tyukavina, A. & Hansen, M.C. 2018. 251: 45–51.
cially in the format of a Talanoa Dialogue Classifying drivers of global forest loss. INPE. Undated. TerraBrasilis [online]. Instituto
(UNFCCC, 2018) – that is, a facilitative Science, 361(6407): 1108–1111. Nacional de Pesquisas Espaciais (INPE).
dialogue between parties to the Paris Ellison, D., Morris, C.E., Locatelli, B., Sheil, [Cited July 2019]. http://terrabrasilis.dpi.
Agreement and non-party stakeholders D., Cohen, J., Murdiyarso, D., et al. 2017. inpe.br
with the goal of finding consensus through Trees, forests and water: cool insights for a IPCC. 2019. Climate change and land: an
transparent, inclusive dialogue. hot world. Global Environmental Change, IPCC special report on climate change,
In 2018, the FWC network also organized 43: 51–61. desertif ication, la nd degra dation,
a session at the International Union of Eriksson, M., Samuelson, L., Jägrud, L., sustainable land management, food security,
Forest Research Organizations (IUFRO) Mattsson, E., Celander, T., Malmer, A., and greenhouse gas fluxes in terrestrial
Joint Conference on Forests and Water in Bengtsson, K., Johansson, O., Schaaf, N., ecosystems. Intergovernmental Panel on
Valdivia, Chile. The aim of the session was Svending, O. & Tengberg, A. 2018. Water, Climate Change (IPCC).
to share progress within the FWC process forests, people: the Swedish experience in Jones, J.A., Wei, X., Archer, E., Bishop,
and to interact with participants on how building resilient landscapes. Environmental K., Blanco, J.A., Ellison, D., Gush, M.,
to promote actions that integrate science, Management, 62: 45–57. https://doi. McNulty, S.G., van Noordwijk, M. &
policy and practice when related to forests org/10.1007/s00267-018-1066-x Creed, I.F. 2019. Forest-water interactions
and water at multiple scales, particularly Estrada-Carmona, N., Hart, A.K., DeClerck, under global change. In: D.F. Levia, D.E.
the landscape level (see the summary of F.A., Harvey, C.A. & Milder, J.C. 2014. Carlyle-Moses, S. Iida, B. Michalzik, K.
the group discussions in Table 2). Here, a Integrated landscape management for Nanko & A. Tischer, eds. Forest-water
wider group of stakeholders participated agriculture, rural livelihoods, and ecosystem interactions. Ecological Studies Series No.
compared with previous meetings, includ- conservation: an assessment of experience [TBD]. Heidelberg, Germany, Springer-
ing participants working in academia, the from Latin America and the Caribbean. Verlag.
private sector and non-profit organizations. Landscape and Urban Planning, 129: 1–11. Kusters, K., Buck, L., de Graaf, M.,
James Dalton acknowledges research Filoso, S., Bezerra, M.O., Weiss, K.C. & Minang, P., van Oosten, C. & Zagt, R.
support from the International Climate Palmer, M.A. 2017. Impacts of forest 2018. Participatory planning, monitoring
Initiative (IKI) of the Federal Ministry restoration on water yield: A systematic and evaluation of multi-stakeholder
for the Environment, Nature Conservation, review. PloS ONE, 12(8): e0183210. platforms in integrated landscape initiatives.
Building and Nuclear Safety (BMUB), García-Martín, M., Bieling, C., Hart, Environmental Management, 62(1): 170–181.
Germany, grant number 13_II_102. James A. & Plieninger, T. 2016. Integrated Laestadius, L., Buckingham, K., Maginnis,
Reed acknowledges research support from landscape initiatives in Europe: multi-sector S. & Saint-Laurent, C. 2015. Before Bonn
IKI BMUB, grant number 18_IV_084. u collaboration in multi-functional landscapes. and beyond: the history and future of forest
Land Use Policy, 58: 43–53. landscape restoration. Unasylva, 66(245): 11.
Hegerl, G.C., Black, E., Allan, R.P., Ingram, Liss Lymer, B., Weinberg, J. & Clausen,
W.J., Polson, D., Trenberth, K.E., et al. T.J. 2018. Water quality management from
2015. Challenges in quantifying changes source to sea: from global commitments
in the global water cycle. Bulletin of to coordinated implementation. Water
References the American Meteorological Society, International, 43(3): 349–360.
96: 1097–1115. https://doi.org/10.1175/ Liu, J., Hull, V., Godfray, H.C.J., Tilman,
Carmenta, R. & Vira, B. 2018. Integration for BAMS-D-13-00212.1 D., Gleick, P., Hoff, H., Pahl-Wostl, C.,
restoration: reflecting on lessons learned from Ilstedt, U., Bargués Tobella, A., Bazié, Xu, Z., Chung, M.G., Sun, J. & Li, S.
the silos of the past. In S. Mansourian & J. H.R., Bayala, J., Verbeeten, E., Nyberg, 2018. Nexus approaches to global sustainable
Parrotta, eds. Forest Landscape Restoration, G., Sanou, J., Benegas, L., Murdiyarso, development. Nature Sustainability, 1(9):
pp. 32–52. London, Routledge. D., Laudon, H., Sheil, D. & Malmer, A. 466.
Creed, I.F. & van Noordwijk, M., eds. 2018. 2016. Intermediate tree cover can maximize Morrison, T.H., Adger, W.N., Brown,
Forest and water on a changing planet: groundwater recharge in the seasonally dry K., Lemos, M.C., Huitema, D., Phelps,
vulnerability, adaptation and governance tropics. Scientific Reports, 6: 21930. J., Evans, L., Cohen, P., Song, A.M.,
opportunities. A global assessment report. Ilstedt, U., Malmer, A., Verbeeten, E. Turner, R., Quinn, T. & Hughes, T.P.
IUFRO World Series, Volume 38. Vienna, & Murdiyarso, D. 2007. The effect of 2019. The black box of power in polycentric
International Union of Forest Research afforestation on water infiltration in the envi ronmental gover nance. Global
Organizations (IUFRO). tropics: a systematic review and meta- Environmental Change, 57: 101934.

Unasylva 251, Vol. 70, 2019/1


45

Nagendra, H. & Ostrom, E. 2012. Polycentric livelihoods in the tropics. Forest Policy and Springgay, E., Dalton, J., Samuelson,
governance of multifunctional forested Economics, 84: 62–71. L., Bernard, A., Buck, A., Cassin, J.,
landscapes. International Journal of the Reid, H., Hou Jones, X., Porras, I., Hicks, Matthews, N., Matthews, J., Tengberg,
Commons, 6(2): 104–133. C., Wicander, S., Seddon, N., Kapos, V., A., Bourgeois, J., Öborn, I. & Reed, J.
Neary, D.G., Ice, G.G. & Jackson, C.R. 2009. Rizvi, A.R. & Roe, D. 2019. Is ecosystem- 2018. Championing the forest–water nexus
Linkages between forest soils and water based adaptation effective? Perceptions – report on the Meeting of Key Forest and
quality and quantity. Forest Ecology and and lessons learned from 13 project Water Stakeholders. Stockholm, Stockholm
Management, 258: 2269–2281. sites. IIED Research Report. London, International Water Institute (SIWI).
O’Connor, A., et al. Forthcoming. The International Institute for Environment and Tengberg, A., Bargues-Tobella, A., Barron,
potential for integration? An assessment Development (IIED). J., Ilstedt, U., Jaramillo, F., Johansson,
of national environment and development Rockström, J., Falkenmark, M., Folke, K., Lannér, J., Petzén, M., Robinson,
policies. C., Lannerstad, M., Barron, J., Enfors, T., Samuelson, L. & Östberg, K. 2018.
Ostrom, E. 2010. Polycentric systems E., Gordon, L., Heinke, J., Hoff, H. & Water for productive and multifunctional
for coping with collective action and Pahl-Wostl, C. 2014. Water resilience for landscapes. Report no. 38. Stockholm,
global environmental change. Global human prosperity. Cambridge, Cambridge Stockholm International Water Institute
Environmental Change, 20(4): 550–557. University Press. (SIWI).
Power, A.G. 2010. Ecosystem services Rockström, J., Falkenmark, M., Karlberg, Tengberg, A. & Valencia, S. 2018. Integrated
and agriculture: tradeoffs and synergies. L., Hoff, H., Rost, S. & Gerten, D. 2009. approaches to natural resources management
Philosophical Transactions of the Royal Future water availability for global food – theory and practice. Land Degradation
Society B: Biological Sciences, 365(1554): production: the potential of green water for and Development, 29: 1845–1857. https://
2959–2971. increasing resilience to global change. Water doi:10.1002/ldr.2946
Reed, J., van Vianen, J., Deakin, E.L., Barlow, Resources Research, 45(7). UNFCCC. 2018. Talanoa Dialogue for
J. & Sunderland, T. 2016. Integrated Seddon, N. 2018. Nature-based solutions: Climate Ambition – synthesis of the
landscape approaches to managing social delivering national-level adaptation and preparator y phase. United Nations
and environmental issues in the tropics: global goals. IIED Briefing. London, Framework Convention on Climate
learning from the past to guide the future. International Institute for Environment and Change (UNFCCC) (available at https://
Global Change Biology, 22(7): 2540–2554. Development (IIED). talanoadialogue.com/presidencies-corner
Reed, J., van Vianen, J., Foli, S., Clendenning, Seddon, N., Turner, B., Berry, P., Chausson, or https://img1.wsimg.com/blobby/go/
J., Yang, K., MacDonald, M., Petrokofsky, A. & Girardin, C.A. 2019. Grounding 9fc76f74-a749-4eec-9a06-5907e013dbc9/
G., Padoch, C. & Sunderland, T. 2017. nature-based climate solutions in sound downloads/1cu4u4230_141793.pdf). u
Trees for life: the ecosystem service biodiversity science. Nature Climate
contribution of trees to food production and Change, 9: 84–87.

Unasylva 251, Vol. 70, 2019/1


46
© R. LINDSAY

Peatlands: the challenge of mapping the world’s


invisible stores of carbon and water
R. Lindsay, A. Ifo, L. Cole, L. Montanarella and M. Nuutinen

W
Peatlands have long been hen Apollo 13 suffered cata- extraordinary exercise in radical thinking
unrecognized or ignored, but they strophic failure during its flight and finite resource management.
will play a crucial role in climate to the Moon in 1970, initially Given the current situation on Spaceship
change and water security and there was confusion and uncertainty. Earth, it is tellingly ironic that the greatest
must be a focus of policy and Commander Jim Lovell spotted a “gas” danger facing the Apollo 13 crew during
research. leaking into space from the Command their remarkable subsequent voyage was
Richard Lindsay is at the Sustainability Module. An hour later, the Command Mod- a buildup of carbon dioxide (CO2) within
Research Institute, University of East London, ule had lost its entire oxygen supply. This their “lifeboat” because the LM’s air fil-
London, United Kingdom of Great Britain and
Northern Ireland.
caused its fuel cells to shut down, leaving ters were unable to process the additional
Averti Ifo is at the Université Marien Ngouabi it without power. If the crew had imme- burden of that gas. Spaceship Earth is also
Ecole Normale Supérieure, Département des diately been able to identify and plug the experiencing dangerous emissions and an
Sciences et Vie de la Terre, Laboratoire de
Géomatique et d’Ecologie Tropicale Appliquée,
leak, the situation need not have become alarming rise in CO2 concentration. As
Brazzaville, the Congo. as critical as it did, but they couldn’t see with Apollo 13, however, even though the
Lydia Cole is at the School of Geography & where the emissions were coming from, buildup of CO2 in the Earth’s atmosphere
Sustainable Development, University of St
Andrews, St Andrews, Scotland, United Kingdom
or why. It became clear that, if they were
of Great Britain and Northern Ireland. to survive, the Lunar Module (LM) must
Luca Montanarella is at the European instead become their lifeboat – although Above: A University of East London research
Commission, Joint Research Centre (JRC),
Ispra, Italy.
the LM was designed to support two men team monitors Plantlife’s Munsary Peatlands
blanket bog, Caithness, northern Scotland,
Maria Nuutinen is Forestry Officer at the for 45 hours, not three men for 90 hours. United Kingdom of Great Britain and
FAO Forestry Department, Rome, Italy. The next four days were to become an Northern Ireland

Unasylva 251, Vol. 70, 2019/1


47

is well documented, the emissions leading land-use change and carbon absorbed by characterize peatlands (Figure 2). Based
to this Earth-bound crisis are proving just terrestrial ecosystems are both subject to on this carbon content, a peat depth of
as difficult to track down. considerable uncertainty (Hansis, Davis only 30 cm contains 327 tonnes of carbon
and Pongratz, 2015). This is because per ha; in comparison, primary tropical
GLOBAL CARBON EMISSIONS – both are extremely difficult to measure rainforest contains 300 tonnes per ha in
ARE WE LOOKING IN THE RIGHT across all the various forms of land-use soil and biomass combined (Blais et al.,
PLACE? intervention and eco­system response. As 2005). This is because the carbon store
The headline figures are simply stated. a pragmatic consequence, the carbon bal- in peat is continuous whereas a forest has
According to the latest data from the Global ance of land-use change, in assessing these gaps between trees – it is said, therefore,
Carbon Project (Le Quéré et al., 2018), global fluxes, has largely been estimated that you can walk through a forest but only
which estimates carbon-flux pathways by quantifying changes in forest cover on a peatland.
based on measured atmospheric values, on the assumption that, compared with Carbon density varies between peat-
the average annual increase in atmospheric the conversion of grasslands to pastures land types, as well as between different
carbon in the period 2008–2017 was 4.7 or croplands, conversion from forest to peatland conditions and even with peat
gigatonnes (Gt). Average annual fossil-fuel open land results in far more significant depth. Generally, the deeper the peat and
emissions in that period were 9.4 Gt of car- losses of both biomass and soil carbon the less disturbed a peatland system, the
bon, and the world’s oceans absorbed some (Houghton, 1999). less dense its carbon content, although
2.5 Gt per year of this. Two other major Although this assumption may hold true this is relative. For example, Warren et al.
pathways contribute to this picture of atmo- for most environments, it is certainly not (2012) recorded a fairly consistent value of
spheric carbon balance: carbon released the case for peatland ecosystems. The larg- around 60 tonnes of carbon per m3 for three
by land-use change, estimated at around est expanses of peatlands occur as open types of Indonesian tropical peatland sys-
1.5 Gt per year, and carbon absorbed by landscapes, and many naturally forested tems ranging in depth from 2.5 m to 12 m,
terrestrial ecosystems, estimated as 3.2 Gt, peatlands have been drained to increase and similar carbon densities can be found
leaving 0.5 Gt unaccounted for (Figure 1). timber production. The World Reference in temperate-zone peat bogs in Scotland
Atmospheric CO 2 concentrations, Base for Soil Resources (WRB) soil clas- possessing several metres of peat in good
fossil-fuel emissions and ocean uptake sification (IUSS Working Group WRB, condition. Even with these lower carbon
are now relatively well documented, but 2015) shows the extraordinary carbon densities, a peat thickness of just 50 cm is
global estimates of carbon emissions from content of the soils (termed histosols) that required at such sites to match the carbon
content of tropical rainforest (compared
with the 30 cm thickness required for thin-
12 ner, denser peat deposits). Moreover, given
the depth of most peatlands (peat depth
10
can extend as much as 60 m below the
8 surface), recent assessments have estimated
that, globally, peatland systems contain
6 an average of 1 375 tonnes of carbon per
Gt carbon

4 ha – more than four times the carbon stored


in an equivalent area of tropical rainforest
2 (Yu et al., 2010; Crump, 2017).
0 Carbon density is one source of varia-
tion, but peat depth gives rise to yet further
-2 levels of uncertainty. The Harmonized
-4 World Soil Database (HWSD) (FAO/
Fossil-fuel Increased Take-up of Ecosystem Land-use IIASA/ISRIC/ISS-CAS/JRC, 2009) takes
emissions carbon carbon by carbon carbon
in the the sea take-up emissions
atmosphere
1
Measured values Estimated values Budget imbalance Estimated annual carbon fluxes to and from
the atmosphere, 2008–2017
Note: Based on documented fossil-fuel emissions, measured atmospheric carbon concentrations, measured carbon
concentrations in the oceans, and the estimated take-up by terrestrial ecosystems and emissions from land-use
activities; 0.5 Gt carbon is unaccounted for in current measurements and estimates.
Source: Global Carbon Project (Le Quéré et al., 2018).

Unasylva 251, Vol. 70, 2019/1


48

2
Organic carbon content for FAO-UNESCO
250.0 soil units to 2 m depth

200.0
and their shallow nature renders them more
amenable to exploitation, degradation and
Organic carbon content (kg/m2)

wholesale loss. Tanneberger et al. (2017)


150.0
sought to produce a harmonized map of
peatlands in Europe based on data presented
100.0 in the first complete review of peatlands
across the continent (Joosten, Tanneberger
and Moen, 2017). Both Tanneberger et al.
50.0 (2017) and Joosten, Tanneberger and Moen
(2017) chose, however, not to specify a
minimum depth of peat for the definition
0.0
of peatland because it was recognized that
different thresholds of peat depth had been
C mb ols
Po ern sols

uv s
rra ls
Ph ley ols
oz ls
uv s

G vis s
yz ls
N ms

is s

s
en ls
eg ls

lo ls
do tz

r s
an ls

Ye ros s
So o s
nc ls
ks
ol m

Fl em
Lu sol

H sol
Po sol

Ve sol

l
ol
Fe iso

ae so

re o

Ar dzo
R so
So so

Pl tiso

Xe so

lo so
An ne

ha
C cris

G ls

e
ds oze
i

ito

to

o
o

applied in different countries, with some


m
A

r
a
h

ignoring thin peat altogether. In the United


Note: Peatland soils are histosols, and many extend to significantly greater depths than 2 m.
Kingdom of Great Britain and Northern
Source: Batjes (1996). Ireland, for example, the figure given by
Tanneberger et al. (2017) for that country’s
1 m depth as its reference depth for each unrecognized (Lindsay, 1993). The soils contribution to the European peat map is
soil unit because many of the national soil that characterize peatlands are hidden 2.6 million ha, but the relevant chapter in
surveys that contribute data to the harmo- below the ground, making it difficult to Joosten, Tanneberger and Moen (2017)
nized database have adopted this threshold. distinguish between peatland and non- gives a figure of 7.4 million ha for “peat
Consequently, the HWSD is severely con- peatland. In addition, the reputation of and peaty soils” (Lindsay and Clough,
strained in its capacity to provide estimates peatlands as unproductive and dangerous 2017). Thus, in the United Kingdom of
of peat depth and carbon storage for the wastelands, good only for conversion to Great Britain and Northern Ireland alone,
global peatland resource. The HWSD is productive uses, has meant that peatlands the area of uncertainty concerning the true
further limited in accurately identifying have also tended to vanish from our col- extent of peatlands amounts to around
the true extent of the global peat resource lective cultural consciousness and so have 4 million ha, almost wholly associated
because of the relatively coarse scale of become more difficult to recognize. Thus, with thin peat. Assuming a depth of 30 cm
mapping and the often small number of peatlands are often labelled as something for this peat, the quantity of carbon stored
field samples used to generate the soil other than peatland, with the result that within this single example of uncertainty in
survey data. Indeed, if there is a consistent their management causes harm that may one nation’s peatland resource approaches
theme running through the underpinning not even be observed. This is dangerous the total estimated annual global emis-
literature of peatland extent and global because the failure to recognize an area sions of 1.5 Gt of carbon resulting from
carbon flux, it is acknowledgement that as a peatland can lead to unexpected and land-use change.
peatland extent and depth are not well sometimes very costly consequences. Such uncertainty is far from unique to
documented, and the land-use changes the United Kingdom of Great Britain and
associated with peatlands are mostly not THIN PEAT – PERIPHERAL BUT Northern Ireland – it is a global issue. What
included in current global atmospheric CRUCIAL does it imply for the extent, condition of,
assessments (Houghton, 1999; Houghton, The issue is particularly crucial for thin- and possible emissions from, the global
2003; Houghton et al., 2012). There are ner peats, essentially those with depths of peatland resource? Yu et al. (2010) gave
many reasons for this, but the underlying 20–60 cm, not only because they tend to widely quoted estimates of 4.4 million km2
cause is that peatlands are “invisible” – cover significantly more area than deep (3 percent of the global land surface) and
both physically and culturally. They have peat but also because they are more easily around 600 Gt of stored carbon for the
been dubbed the “Cinderella habitat” confused with other habitats and more known extent of the global peat resource,
because they provide so many ecosys- easily destroyed. Thin peat deposits are based largely on documented areas of deep
tem services yet continue to go largely consequently more challenging to map, peat. These estimates alone mean that the

Unasylva 251, Vol. 70, 2019/1


49

© R. LINDSAY
One hectare of peat only 30 cm deep holds as much carbon as 1 hectare of primary tropical rainforest, yet it may be mistaken for other habitats
such as heathland and so managed inappropriately. Such a thin layer of peat is more easily destroyed by inappropriate management – the
single pass of a plough, for example – than is the case for the loss of the carbon store held in a tropical rainforest, where, even after felling
and burning, the roots and stumps of the forest remain. The loss of thin peat does not attract as much world attention as the loss of tropical
rainforest, however

known global peatland resource contains resulting from destruction through lack of die, their remains accumulate in situ
more carbon than all the world’s vegetation awareness, or positively, by halting emis- because waterlogging slows decomposition
combined (Scharlemann et al., 2014). The sions, preserving the carbon, bringing back to such an extent that a proportion of this
fact that even thin peats (i.e. those peats other ecosystem services and, eventually, plant material and its associated carbon
most vulnerable to land-use change) have over longer timescales, restoring the sys- is preserved in what becomes peatland,
the potential to release as much carbon per tems once more to carbon sinks. often on millennial timescales. Its water-
unit area as the clearing of primary tropi- logged state means that peat is commonly
cal forest lends particular urgency to the THE CONSEQUENCES OF PEATLAND as much as 95 percent water by weight
need for accurate mapping of these mostly MISMANAGEMENT and 85 percent by volume, meaning that
overlooked but potentially very large areas peatlands are significant contributors to
of thinner peat. Even small changes in the The release of carbon water control, often at the landscape scale.
mapped extent of national and global peat Peatlands are wetlands of major sig- The general land-use trend for these wet
resources could mean substantial changes nificance in terms of carbon storage and landscapes, however, has been to drain
to the picture of associated carbon fluxes – release because waterlogging preserves them in order to make them more amen­
whether negatively, in terms of emissions dead plant matter. When wetland plants able to exploitation (IPBES, 2018). When

Unasylva 251, Vol. 70, 2019/1


50

water is removed from the peat matrix as of Great Britain and Northern Ireland, the also have major implications for carbon
a result of drainage, the peat undergoes area of East Anglia known as the Fens emissions (Page et al., 2002) could now
significant shrinkage through “primary once consisted of a peatland covering be stimulating greater interest in establish-
consolidation” and “secondary compres- 1 500 km2. Records from the seventeenth ing precisely where the peatlands are and
sion”, resulting in subsidence of the ground century indicate that this accumulated how best to manage them. In recent years,
surface. Moreover, when air penetrates the peat was a key factor in holding back the several substantial peatland systems have
normally waterlogged peat it initiates rapid sea from this large drainage basin (Darby, been reclassified as peatland, having previ-
decomposition and the release of long-term 1956, p. 107). The wholesale drainage of ously been described as other habitat types.
carbon into the atmosphere (“oxidative the area in the eighteenth and nineteenth
wastage”), giving rise to carbon emissions centuries by “adventurers” (whom today REGIONAL STATUS OF PEATLAND
as well as further ground subsidence. might be called financial speculators) to MAPPING FOR CARBON, WATER
It is unfortunate, therefore, that the grow arable crops on the rich peat soil has AND BIODIVERSITY
drainage of such systems is inadequately since given rise to some of England’s finest Substantial progress has been made in
captured in the present global atmo- agricultural land. There has been a signifi- peatland mapping and the development
spheric model of carbon fluxes in terms cant price to pay, however, beyond the loss of policy processes in the last decade or
of emissions due to land-use change in of the area’s formerly rich biodiversity. The so, as illustrated by the examples below.
peatlands (Houghton et al., 2012). Such peat soils subject to intensive agriculture Nevertheless, there are likely many more
emissions could be significantly larger than release as much as 8 tonnes of carbon per areas of overlooked peatlands awaiting dis-
shown in Figure 1 but in that case they ha annually through oxidative wastage covery, particularly in Africa but also areas
must also be balanced by greater carbon (Evans et al., 2016), and the ground surface of thin peat on every continent currently
capture than indicated, resulting in the has subsided to such an extent that many classed as something other than peatland.
same overall rise in atmospheric CO2. areas are now as much as 3 m below sea
Should this additional take-up of CO2 by level. Continued farming is only possible The Congo’s vast peatlands
terrestrial ecosystems begin to fail as a because of substantial and very expensive Deep in the Congo Basin, in an area that is
result of climate change, however, emis- drainage infrastructure, and the cost is now enormously difficult to access, a peat-bog
sions from land-use change could take so high, and the threat of rising sea levels system was brought to light only recently
on considerable added importance. The and subsiding ground levels so serious, by scientific collaboration among sev-
main alternative source of estimates for that the country’s Environment Agency is eral teams of researchers. This peatland
emissions due to land-use change are the discussing the need to move entire com- complex is now recognized as the largest
data collated from the individual national munities to safer ground in the foreseeable known continuous peat-bog system in the
greenhouse-gas accounting reports sub- future (UK Environment Agency, 2019). tropics, at almost 145 000 km2, two-thirds
mitted under the Kyoto Protocol. The Similar issues are being discussed in of which is in the Democratic Republic of
guidance provided to those assembling coastal areas of Southeast Asia, where the Congo and the remaining one-third in
these national reports (Intergovernmental extensive peatlands have been converted the eastern part of the Congo (Dargie et
Panel on Climate Change, 2014) has wid- to major rice projects and, more recently, to al., 2017). The area is so enormous that it
ened to include procedures for estimating oil-palm and acacia plantations; this has led encompasses two very large Ramsar sites,
carbon emissions from peatland systems to widespread peatland fires, and peatland Lac Télé in the Congo and Ngiri-Tumba-
subject to, for example, drainage for agri- subsidence is in danger of causing huge Maindombe in the Democratic Republic
cultural purposes. Even the collation of areas of coastal flooding (Hooijer, 2012). of the Congo, the latter being the world’s
this information provides only a partial These and other problems have arisen time second-largest Ramsar site. The known
picture, however, because some nations and time again, either because there was extent of the newly identified peatland
do not participate and all nations have a failure to recognize that an area was a area amounts to almost 4 percent of the
difficulty in deciding the area over which peatland or because the consequences of Congo Basin (the world’s second-largest
the particular peat-related emission factors exploiting the peatland were insufficiently river basin). With measured peat depths
should be applied because the extent of understood. Both these reasons continue of 0.3–5.9 m, the recorded peatland area
peatlands is so poorly known. to represent major challenges worldwide, is estimated to contain 30 Gt of carbon;
and even major deposits of deep peat have thus, this peatland system contains nearly
The problem of subsidence continued to be overlooked, misclassified 5 percent of the carbon contained in the
Peat subsidence itself gives rise to undesir- or subsumed under some other habitat type world’s known peatlands. This peatland
able consequences beyond those of carbon (as explored below). On the other hand, plays an essential role in the regional
loss. In the lowlands of the United Kingdom growing recognition that such actions climate of the Congo Basin and makes a

Unasylva 251, Vol. 70, 2019/1


51

significant and active contribution to the


catchment dynamics of the Congo River,
which is second only to the Amazon in
the volume of its discharge. The peatland
complex constitutes a huge reservoir of
freshwater and, because it so large that it
often covers entire interfluves, it is a key
water source for various tributary systems
(e.g. the Oubangui and Sangha) that flow
through this vast ecological zone.
Driven by concerns about the potential
impacts of climate change in the region,
researchers in the CongoPeat project are
seeking to understand how the peatlands
originally developed and what has main-
tained them as waterlogged, peat-forming
systems for the past 10 000 years or so,
thereby enabling the establishment of the
area’s exceptional biodiversity. In addi-
tion to preparing preliminary maps of the
peatlands to enable improved land-use
planning, the CongoPeat team is attempt-
ing to understand the water balance of
these systems because the majority appear
to be water-shedding, meaning they rely
solely on direct precipitation inputs for
their water supply (i.e. they are ombrotro-
phic bogs). In such systems, losses from
evaporation and drainage by gravity flow
must be balanced by precipitation inputs,
and there may be significant consequences
if these inputs and outputs are altered by a
regional decline of rainfall or longer-term
climate change.
Given that the Congo peatlands rely
on the basin’s overall rainfall pattern,
it is significant that recent recorded
data and publications on rainfall in the
(Republic of the) Congo have shown a
marked decline in rain inputs. This is
probably partly due to deforestation but
mainly to recent negative trends in atmo-
spheric and oceanic parameters: that is,
the Atlantic Multi-Decadal Oscillation, the
North Atlantic Oscillation and the Southern
© L. COLE

Oscillation Index (Ibiassi Mahoungou


et al., 2017; Ibiassi Mahoungou, 2018). Local people receive training on the use
Particularly in light of these trends, impor- lost through evaporation, evapotranspira- of a mobile-phone-based application
for collecting information on Mauritia
tant questions need to be answered: How tion and lateral drainage? flexuosa productivity in a palm swamp
much rainfall is required to maintain satu- In addition to studies aimed at determin- (regionally known as an aguajal) in the
rated conditions? And how much water is ing the water balance, field surveys have PMFB, western Amazonia

Unasylva 251, Vol. 70, 2019/1


52

revealed the exceptional biodiversity of


these peatlands, including iconic species
such as the forest elephant and hippopota-
mus. The three large African primates
– gorillas, chimpanzees and bonobos – all
have significant populations there (Fay et
al., 1989; Fay and Agnagna, 1992; Blake
et al., 1995), and the region supports more
than 350 bird species, including a number
of endemic species (Evans and Fishpool,
2001). This highlights the fact that, because
peatlands are so often overlooked, the
remarkable and often highly distinctive
biodiversity they support also remains
hidden or is assumed to be dependent on
other habitat types whereas peatlands may
actually constitute the core habitat areas
for certain key species (e.g. Singleton and
van Schaik, 2001; Baker et al., 2010).
Resources spent on maintaining habitats
assumed to be vital for this biodiversity
may be wasted if the true core habitat
features are lost in the meantime through
misplaced actions.

Peatlands in the Amazon


A similar story of discovery has unfolded in
the world’s largest river basin, the Amazon,
in the last few decades. Some of the first
published studies on the peatlands of the
Pastaza-Marañón Foreland Basin (PMFB)
in western Amazonia in the northern
Peruvian lowlands described a peat-rich
area of approximately 100 000 km2 con-
taining 2–20 Gt of carbon (Lähteenoja
et al., 2009). Since then, research has
been carried out to refine these estimates
and to understand more about the area’s
developmental processes (Roucoux et al.,
2013; Kelly et al., 2017) and eco­system
characteristics (Draper et al., 2014, 2018).
Understanding the interannual flood
variability, associated environmental
disturbances and river-channel dynamics
of the Amazon Basin is key to under-
standing the development of its peatlands
© L. COLE

(Gumbricht et al., 2017). Such factors have


created a complex arrangement of environ- A palm swamp (aguajal)
dominated by Mauritia flexuosa
ments that are waterlogged throughout the Unlike in Southeast Asia, where coastal in the PMFB, western Amazonia
year and thus ideal for the development of domes are the dominant form in which
peatlands (e.g. Householder et al., 2012). peat is found (Dommain, Couwenberg and

Unasylva 251, Vol. 70, 2019/1


53

© R. LINDSAY
The carnivorous sundew Drosera binnata
on the margin of a peat pool formed within
a patterned fen peatland near Moon Point, Locally, peatlands are often referred to as interannual flooding variability of the
Fraser Island, Australia “sucking” environments (chupaderas in basin’s rivers, with waters rising in some
Spanish), illustrating the lived experience places by up to 10 m, means that draining
Joosten, 2011), many of the PMFB’s peat- of traversing them. the peatlands would be near-impossible.
lands are small, discrete and transient over Although large and significant, the The lack of a coherent road network
geological timescales. To date, depths of up PMFB is just one of the basins in the also prevents the overland transportation
to 7.5 m have been found (Lähteenoja et al., Amazon that contains peat. Others have of machinery and human resources to
2009), covered by vegetation communities been classified in the eastern Amazon in support industrial-scale drainage. Plans to
varying from open grass and sedge-rich Peru (Householder et al., 2012) and in the greatly extend the regional infra­structure
ecosystems to pole forests and palm Brazilian Amazon (Lähteenoja, Flores and enhance extractive capabilities in
swamps, where one particular palm of and Nelson, 2013), and there are probably the future, however, would increase the
economic value, Mauritia flexuosa, com- many more, currently “invisible” areas vulnerability of the peatlands of the
monly dominates (Lähteenoja et al., 2009). that need to be formally identified and PM F B a nd beyond ( Roucoux et
People living in and around the peatlands classified and which are subject to various al., 2017). The challenge for the
of the PMFB classify and use these ecosys- threats. Compared with the situation in scientific community is to evalu-
tems in various ways (Schulz et al., 2019), Southeast Asia (Page and Hooijer, 2016), ate the contributions that Amazonian
although they tend to avoid them when many of the Amazon’s peatlands are rela- peatlands make to carbon and water
alternative landscape types are available tively intact and under limited immediate cycl i ng, t hought to be of huge
(L. Cole, personal communication, 2019). threat of drainage or conversion. The significance on a local to global scale

Unasylva 251, Vol. 70, 2019/1


54

(Gumbricht et al., 2017), before such con- Despite the continued efforts of the it should be possible to find at least
tributions are compromised. European Union member states and policy- some places where the rod or stick can
makers to reverse the trend and protect and be made to penetrate to a depth of at
Fraser Island’s newly discovered restore peatlands and other wetlands and least 30 cm with relative ease.
peatland avoid their continued drainage and degra- 2. Take a sample from a depth of
On Fraser Island off the coast of the dation, little research exists on the direct 20–30 cm, air-dry the sample and see
Australian state of Queensland, areas for- effectiveness or cross-sectoral impacts of if it will burn. The high organic matter
merly classified as relatively uninteresting the numerous interventions. The European content of peat means that, once dry,
“wet heath” have now been acknowledged Union’s environmental laws and incentive it should ignite readily.1
as highly distinctive peatland systems that schemes, particularly those linked to the Perhaps the greatest challenge in
support a significant number of endangered Natura 2000 framework, have established determining the true extent of peatlands
species (Fairfax and Lindsay, forthcom- a strong protection regime for peatlands, identified through surveys is the resolution
ing). Having previously been excluded but other legislative frameworks, including used. If a small pocket of peat measuring
from the Fraser Island World Heritage Site, the Common Agricultural Policy and the 100 m × 100 m (i.e. 1 ha) × 30 cm deep can
these peatlands may now be incorporated renewable-energy policy, have arguably contain as much carbon as the same area of
in it as important ecosystem components. yielded opposite effects by providing primary tropical rainforest, there is evident
With sympathetic management, the peat- perverse incentives. The specific effects benefit in ensuring that the mapping resolu-
lands also have the potential to be valuable of the European Union’s climate policy tion is sufficiently fine to identify areas
carbon sinks and key hinterland providers frameworks on peatlands have not yet been of this size. Ideally, therefore, mapping
of iron-rich waters to support the role of fully addressed (Peters and von Unger, would be undertaken at a scale of 1:10 000,
coastal mangroves as nursery grounds for 2017). A new effort may be initiated, how- but for large areas a scale of 1:20 000
local fish populations. ever, in response to a recent resolution by may be the highest resolution achievable
the United Nations Environment Assembly with current technology and available
Peatlands in Europe (2019), which calls for more emphasis on resources.
Tanneberger et al. (2017) estimated the the conservation, sustainable management
area of peatland in Europe at 593 727 km². and restoration of peatlands worldwide, Recommendations for policymakers
Mires, which by definition are dominated as also recommended in a recent assess- Policymakers should:
by living and peat-forming plants, were ment by the Intergovernmental Platform • Verify whether it is likely that more
found to cover more than 320 000 km² for Biodiversity and Ecosystem Services peatlands would be found in the
(around 54 percent of the total peatland (IPBES, 2018). country.
area). If shallow peatlands (< 30 cm • Prioritize the mapping of peatlands at
peat) in the European part of the Russian CONCLUSIONS a scale of at least 1:20 000 but ideally
Federation are included, the total peat- Here, we make some simple recommenda- 1:10 000.
land area in Europe is more than 1 million tions for improving understanding of the • Map past, ongoing and planned man-
km2 – almost 10 percent of the total land true extent of peatlands on the planet – the agement (“activity data”, under the
area. Peatlands are distributed widely first step in their protection, for all the United Nations Framework Conven-
among the European Union countries, benefits this will bring. tion on Climate Change), including
with concentrations in northern, central existing drainage infrastructure and
and eastern Europe (Germany, Ireland, the Recommendations for action – finding other livelihood activities in the area
Netherlands, Poland, the United Kingdom the peat (e.g. fishing and peat extraction).
of Great Britain and Northern Ireland, and Two simple steps can be used to determine • Include peatland maps in planning pro-
the Nordic and Baltic countries). Official whether you are standing on peat: cesses from the local to the regional
policy research efforts, political appraisals 1. Peat is a relatively soft soil, so it should scale, not only for climate and bio-
and firm legislative provisions exist that be possible to push a rod or stick with diversity benefits but also for water
recognize the need to protect peatlands a diameter of 6–8 mm at least 30 cm security and disaster risk reduction.
and the inherent vulnerability of their into the soil using only hand pressure. • Protect undrained peatlands to avoid
soils. In practice, however, the degrada- We use a length of 6-mm-diameter activities that might cause important
tion of these ecosystems is continuing threaded steel rod – widely available
1
Note that soils containing agrochemical
across the European Union, due mainly around the world. This may not work
residues can release noxious or toxic
to drainage for agriculture and forestry and so easily in some tropical peats that fumes when heated. Please take suitable
peat extraction for fuel and horticulture. consist largely of wood but, even so, precautions.

Unasylva 251, Vol. 70, 2019/1


55

changes to their hydrology and associ- The next generation looks to us to sequestration of tropical peat domes in south-
ated ecosystem services. address the challenge of climate change east Asia: links to post-glacial sea-level
• Budget for the restoration of drained so that they, and Spaceship Earth, can changes and Holocene climate variability.
peatlands and for documenting and survive. Because, for them, there is no Quaternary Science Reviews, 30(7–8):
developing drainage-free livelihood LM, there is no lifeboat, there is no 999–1010.
options. alternative. u Draper, F.C., Honorio Coronado, E.N.,
• If peatland drainage continues, invest Roucoux, K.H., Lawson, I.T., Pitman,
in the development of systems for fire N.C.A., Fine, P.V.A., Phillips, O.L., Torres
risk assessment, fire reduction and fire Montenegro, L.A., Valderrama Sandoval,
management. E., Mesones, I., García-Villacorta, R.,
• Harmonize incentives, laws and law Ramirez Arévalo, F.R. & Baker, T.R.
enforcement to support these goals. References 2018. Peatland forests are the least diverse
• Communicate to all decision-makers, tree communities documented in Amazonia,
stakeholders and the public the impor- Baker, J.R., Whelan, R.J., Evans, L., Moore, but contribute to high regional beta-diversity.
tance of peatlands for water, biodiver- S. & Norton, M. 2010. Managing the ground Ecography, 41, 1256–1269.
sity and climate change. parrot in its fiery habitat in south-eastern Draper, F.C., Roucoux, K.H., Lawson, I.T.,
• Monitor the status of peatlands to Australia. Emu: Austral Ornithology, 110(4): Mitchard, E.T.A., Honorio Coronado,
detect potential signs of emerging 279–284. E.N., Lähteenoja, O., Torres Montenegro,
drainage-based land uses and land- Batjes, N.H. 1996. Total carbon and nitrogen L., Valderrama Sandoval, E., Zaráte, R.
use impacts. in the soils of the world. European Journal & Baker, T.R. 2014. The distribution and
• Report on the status of peatlands of Soil Science, 47: 151–163. amount of carbon in the largest peatland
against the Sustainable Develop- Blais, A.M., Lorrain, S., Plourde, Y. & complex in Amazonia. Environmental
ment Goals and under international Varfalvy, L. 2005 Organic carbon densities Research Letters, 9(12).
conventions. of soils and vegetation of tropical, temperate Evans, C., Morrison, R., Burden, A.,
and boreal forests. In: A. Tremblay, L. Williamson, J., Baird, A., Brown, E.,
Final thoughts Varfalvy, C. Roehm & M. Garneau, eds. Callaghan, N. et al. 2016. Lowland peatland
Spaceship Earth is not a new concept, Greenhouse gas emissions — fluxes and systems in England and Wales – evaluating
but, around the globe, young people’s processes, pp. 155–185. Environmental greenhouse gas fluxes and carbon balances.
active responses to the climate protest of Science. Berlin, Germany, Springer. Final report to DEFRA on Project SP1210.
schoolgirl Greta Thunberg suggest that the Blake, S., Rogers, E., Fay, J.M., Ngangoué, Bangor, UK, Centre for Ecology and
youth of today perhaps grasp the reality M. & Ebeke, G. 1995. Swamp gorillas in Hydrology.
of this concept rather more urgently than northern Congo. African Journal of Ecology, Evans, M.I. & Fishpool, L.D.C. 2001.
have preceding generations. Young people 33, 285–290. Important bird areas in Africa and associated
are looking to those in power to make Crump, J., ed. 2017. Smoke on water – islands: priority sites for conservation.
the same kinds of bold and imaginative countering global threats from peatland loss Cambridge, UK, Birdlife International &
decisions as the highly focused team who and degradation. A UNEP rapid response Pisces Publications.
brought the Apollo 13 crew safely back to assessment. Nairobi, Kenya, United Nations Fairfax, R. & Lindsay, R. Forthcoming. An
Earth. Identifying the true extent of the Environment Programme (UNEP), and overview of the patterned fens of Great Sandy
world’s peatlands and working to return Arendal, Norway, GRID-Arendal. www. Region, far eastern Australia. Mires and Peat.
them to sinks rather than sources of carbon grida.no FAO/IIASA/ISRIC/ISS-CAS/JRC. 2009.
is undoubtedly a difficult challenge. But, Darby, H.C. 1956. The draining of the Fens. Harmonized World Soil Database (version
in the words of the late John F. Kennedy Second edition. Cambridge, Cambridge 1.1). Rome, FAO and Laxenbrug, Austria,
(the 35th president of the United States of University Press. International Institute for Applied Systems
America and a leading proponent of the Dargie, G.C., Lewis, S.L., Lawson, I.T., Analysis (IIASA).
United States of America’s Space Program Mitchard, E.T.A., Page, S.E., Bocko, Y.E. Fay, J.M. & Agnagna, M. 1992. Census of
in the 1960s), we choose to do these things & Ifo, S.A. 2017. Age, extent and carbon gorillas in northern Republic of Congo.
“not because they are easy, but because storage of the central Congo Basin peatland American Journal of Primatology, 27,
they are hard, because that goal will serve complex. Nature, 542: 86–90. DOI: 10.1038/ 275–284.
to organize and measure the best of our nature21048 Fay, J.M., Agnagna, M., Moore , J. & Oko,
energies and skills, because that challenge Dommain, R., Couwenberg, J. & Joosten, R. 1989. Gorillas (Gorilla g. gorilla) in the
is one that we are willing to accept”. H. 2011. Development and ca rbon Likouala swamp forests of North Central

Unasylva 251, Vol. 70, 2019/1


56

Congo: preliminary data on populations IPBES. 2018. The IPBES assessment report – Status, distribution and conservation.
and ecology. International Journal of on land degradation and restoration. L. Stuttgart, Germany, Schweitzerbart Science
Primatology, 10, 477–486. Montanarella, R. Scholes & A. Brainich, Publishers.
Gumbricht, T., Roman-Cuesta, R.M., eds. Bonn, Germany, Secretariat of Page, S.E. & Hooijer, A. 2016. In the line
Verchot, L.V., Herold, M., Wittmann, the Intergovernmental Science-Policy of fire: the peatlands of Southeast Asia.
F.K., Householder, E., Herold, N. & Platform on Biodiversity and Ecosystem Philosophical Transactions of the Royal
Murdiyarso, D. 2017. An expert system Services (IPBES). https://doi.org/10.5281/ Society B: Biological Sciences, 371.
model for mapping tropical wetlands and zenodo.3237392 Page, S.E., Siegert, F., Rieley, J.O., Boehm,
peatlands reveals South America as the Intergovernmental Panel on Climate H-D.V., Jaya, A. & Limin, S. 2002. The
largest contributor. Global Change Biology, Change (IPCC). 2014. 2013 Supplement amount of carbon released from peat and
23(9): 3581–3599. to the 2006 IPCC Guidelines for National forest fires in Indonesia during 1997. Nature,
Hansis, E., Davis, S.J. & Pongratz, J. 2015. Greenhouse Gas Inventories: Wetlands. T. 420: 61–65.
Relevance of methodological choices for Hiraishi, T. Krug, K. Tanabe, N. Srivastava, Peters, J. & von Unger, M. 2017. Peatlands
accounting of land use change carbon J. Baasansuren, M. Fukuda & T.G. Troxler, in the EU regulatory environment. BfN-
fluxes. Global Biogeochemical Cycles, 29: eds. Switzerland, Intergovernmental Panel Skripten 454. DOI 10.19217/skr454, Bonn,
1230–1246. on Climate Change (IPCC). Germany.
Hooijer, A., Page, S., Jauhiainen, J., Lee, IUSS Working Group WRB. 2015. World Roucoux, K.H., Lawson, I.T., Baker, T.R.,
W.A., Lu, X.X., Idris, A. & Anshari G. Reference Base for Soil Resources 2014, Del Castillo Torres, D., Draper, F.C.,
2012. Subsidence and carbon loss in drained update 2015. International soil classification Lähteenoja, O., et al. 2017. Threats to intact
tropical peatlands. Biogeosciences, 9: 1053– system for naming soils and creating legends tropical peatlands and opportunities for their
1071. for soil maps. World Soil Resources Reports conservation. Conservation Biology, 31(6):
Houghton, R.A. 1999. The annual net flux of No. 106. Rome, FAO. 1283–1292.
carbon to the atmosphere from changes in Joosten, H., Tanneberger, F. & Moen, A., eds. Roucoux, K.H., Lawson, I.T., Jones,
land use 1850–1990. Tellus, 51B: 298–313. 2017. Mires and peatlands of Europe – status, T.D., Baker, T.R., Honorio Coronado,
Houghton, R.A. 2003. Revised estimates distribution and conservation. Stuttgart, E.D., Gosling, W.D. & Lähteenoja,
of the annual net flux of carbon to the Germany, Schweitzerbart Science Publishers. O. 2013. Vegetation development in an
atmosphere from changes in land use and Kelly, T., Lawson, I.T., Roucoux, K.H., Baker, Amazonian peatland. Palaeogeography,
land management 1850–2000. Tellus, 55B: T.R., Jones, T.D. & Sanderson, N.K. 2017. Palaeoclimatology, Palaeoecology, 374:
378–390. The vegetation history of an Amazonian 242–255.
Houghton, R.A., House, J.I., Pongratz, J., domed peatland. Palaeogeography, Scharlemann, J.P.W., Tanner, E.V. J.,
van der Werf, G.R., DeFries, R.S., Hansen, Palaeoclimatology, Palaeoecology, 468: Hiederer, R. & Kapos, V. 2014. Global
M.C., Le Quéré, C. & Ramankutty, N. 129–141. soil carbon: understanding and managing
2012. Carbon emissions from land use Lähteenoja, O., Flores, B.M. & Nelson, B.W. the largest terrestrial carbon pool. Carbon
and land-cover change. Biogeosciences, 9: 2013. Tropical peat accumulation in Central Management, 5(1): 81–91.
5125–5142. Amazonia. Wetlands, 33: 495–503. Schulz, C., Martín Brañas, M., Núñez Pérez,
Householder, J.E., Janovec, J.P., Tobler, Lähteenoja, O., Ruokolainen, K., Schulman, C., Del Aguila Villacorta, M., Laurie,
M.W., Page, S. & Lahteenoja, O. 2012. L. & Oinonen, M. 2009. Amazonian N., Lawson, I.T. & Roucoux, K.H. 2019.
Peatlands of the Madre de Dios River of peatlands: an ignored C sink and potential Peatland and wetland ecosystems in Peruvian
Peru: distribution, geomorphology, and source. Global Change Biology, 15(9): Amazonia: indigenous classifications and
habitat diversity. Wetlands, 15: 2311–2320. 2311–2320. perspectives. Ecology and Society, 24(2): 12.
Ibiassi Mahoungou, G. 2018. Variabilité Le Quéré, C., Andrew, R.M., Friedlingstein, Singleton, I. & van Schaik, C.P. 2001.
pluviométrique en République du Congo: P., Sitch, S., Hauck, J., Pongratz, J., Orangutan home range size and its
dynamique océanique et atmosphérique. Pickers, P.A., Korsbakken, J.I., et al. 2018. determinants in a Sumatran swamp forest.
Editions Universitaires Européennes. Global carbon budget 2018. Earth System International Journal of Primatology, 22(6):
Ibiassi Mahoungou, G., Pandi, A. & Science Data, 10: 2141–2194. 877–911.
Ayissou L. 2017. Extrêmes hydrologiques Lindsay, R.A. 1993. Peatland conservation – Tanneberger, F., Tegetmeyer, C., Busse,
et variabilité décennale des précipitations from cinders to Cinderella. Biodiversity and S., Barthelmes, A., Shumka, S., Moles
saisonnières dans le bassin versant du fleuve Conservation, 2: 528–540. Mariné, A., et al. 2017. The peatland map
Congo à Brazzaville (République du Congo). Lindsay, R.A. & Clough, J. 2017. United of Europe. Mires and Peat, 19(22): 1–17
In: B.M. Mengho, ed. Géographie du Congo, Kingdom. In: H. Joosten, F. Tanneberger & (available at www.mires-and-peat.net/pages/
pp. 207–223. Paris, L’Harmattan. A. Moen, eds. Mires and peatlands of Europe volumes/map19/map1922.php).

Unasylva 251, Vol. 70, 2019/1


57

UK Environment Agency. 2019. Draft National United Nations Environment Assembly. 2019. method to assess carbon stocks in tropical
Flood and Coastal Erosion Risk Management Conservation and sustainable management of peat soil. Biogeosciences, 9: 4477–4485.
Strategy for England. Rotherham, UK peatlands. UNEP/EA.4/L.19, Nairobi, 11–15 Yu, Z., Loisel, J., Brosseau, D.P., Beilman,
(available at https://consult.environment- March 2019. D.W. & Hunt, S.J. 2010. Global peatland
agency.gov.uk/fcrm/national-strategy-public/ Warren, M.W., Kauffman, J.B., Murdiyarso, dynamics since the last glacial maximum.
user_uploads/fcrm-strategy-draft-final-1- D., Anshari, G., Hergoualc’h, K., Geophysical Research Letters, 37: L13402. u
may-v0.13-as-accessible-as-possible.pdf). Kurnianto, S., et al. 2012. A cost-efficient

Unasylva 251, Vol. 70, 2019/1


58

Fire, forests and city water supplies


D.W. Hallema, A.M. Kinoshita, D.A. Martin, F.-N. Robinne, M. Galleguillos,
S.G. McNulty, G. Sun, K.K. Singh, R.S. Mordecai and P.F. Moore

F
The changing role of fire in forest orest landscapes generate 57 percent
landscapes shows that strategic of runoff worldwide and supply
forest management is necessary to water to more than 4 billion peo-
safeguard urban water supplies. ple (Millennium Ecosystem Assessment,
2005). As the world population continues
Dennis W. Hallema (dwhallem@ncsu.edu) and Mauricio Galleguillos is Assistant Professor at to increase, there is a strong need to under-
Ge Sun are Research Hydrologists at the United Universidad de Chile, Santiago, Chile. stand how forest processes link together
States Department of Agriculture (USDA) Steven G. McNulty is Director of the USDA
Forest Service Southern Research Station, North Southeast Regional Climate Hub, North in a cascade to provide people with water
Carolina, United States of America. Carolina, United States of America. services like hydropower, aquaculture,
Alicia M. Kinoshita is Associate Professor of Kunwar K. Singh is Research Fellow at North drinking water and flood protection (Car-
Water Resources Engineering at San Diego State Carolina State University, Raleigh, North
University, California, United States of America. Carolina, United States of America. valho-Santos, Honrado and Hein, 2014).
Deborah A. Martin is Research Hydrologist Rua S. Mordecai is Science Coordinator at the
Emeritus at the US Geological Survey (retd.), United States Fish and Wildlife Service, Raleigh,
Boulder, Colorado, United States of America. North Carolina, United States of America. Controlled burns promote forest health by
François-Nicolas Robinne is Postdoctoral Peter F. Moore is Forestry Officer at the FAO cleaning up fuels and promoting tree growth,
Fellow at the University of Alberta, Edmonton, Forestry Department, Rome, Italy. with indirect benefits for the quality of forest
Alberta, Canada. water resources
© DENNIS HALLEMA, USDA FOREST SERVICE SOUTHERN RESEARCH STATION

Unasylva 251, Vol. 70, 2019/1


59

1
Global wildfire risk to water security
based on fire activity, vegetation,
geography, water availability and
socio-economic development

Note: Wildfire risk to water security is shown on a scale from 0 (minimum risk) to 100 (theoretical maximum risk potential).
Source: Robinne et al. (2018), used here under a CC BY 4.0 licence.

Wildfire is a major disturbance affecting are conducive for another fire. less water immediately after fire and, in
forested watersheds and the water they Extreme and hazardous wildfires, on the environments influenced by snow, more
provide (Box 1) (Paton et al., 2015). Several other hand, can cause erosion, gullying, snow can accumulate in forest clearings
regions have experienced shifts in wildfires soil loss and flooding – and, in severe (Kinoshita and Hogue, 2015; Hallema et
from natural ignition sources (primarily cases, even debris flows and flash floods al., 2019). Therefore, accounting for wild-
lightning) to ignitions dominated by human – by removing the protective functions fire impacts on forests in water planning
activities, especially in areas where popu- of forests on hillsides (Ebel and Moody, has become a priority for the nexus of fire,
lations are increasing (Moritz et al., 2014; 2017). Extreme wildfires have become water and society or, in other words, the
Balch et al., 2017). Occasional wildfire is more common after decades of fire sup- connection between fire risk and water
essential for the health and functioning of pression, allowing forests to become much security (Figure 1) (Martin, 2016). In this
fire-adapted ecosystems through its effects denser with vegetation and causing more article, we discuss managed forest land-
on nutrient cycling, plant diversity and fuels to build up over time. Combined with scapes as nature-based solutions for water
succession, and pest regulation (Pausas increasing summer drought, this can have and explore how fire affects the provision
and Keeley, 2019). It also reduces the risk impacts on water yield and the ability of of water-related services.
of subsequent wildfires until a forest has upstream forests to deliver high-quality
accumulated sufficient fuels and conditions water because forest vegetation uses WATER SERVICES FROM FORESTS
In many areas, swimming in a river,
Box 1 preparing food and irrigating the garden
Key facts on fire and forest water resources have a commonality: they rely on water
services provided by upstream forests (Sun,
• Globally, an average of 400 million ha of land was burned annually in the period 2003–2016, Hallema and Asbjornsen, 2018). Water
of which an estimated 19 million ha per year was forest (Melchiorre and Boschetti, 2018). ecosystem services, also called hydrologic
• Tropical forests represent the largest proportion of forested area burned (65.9 percent services, provide a range of direct and
between 2003 and 2016) (Melchiorre and Boschetti, 2018). indirect benefits and associated values.
• Wildfires in the United States of America result in up to 10 percent more surface water Most forest hydrologic services – such
annually – and 10–50 percent more in regions with severe wildfires (Hallema et al., 2019; as hydropower generation, power plant
Kinoshita and Hogue, 2015). cooling, irrigation, aquaculture and flood
• Ninety percent of the world’s cities with populations larger than 750 000 use water from mitigation – can be expressed in terms of
forested watersheds, yet nine out of ten of these watersheds show signs of water-quality a market value. Some services, however,
degradation (McDonald et al., 2016). have intrinsic, non-market values, such as
• Controlled burns (also called prescribed fires) clean up dead vegetation and reduce the aquatic ecosystem quality and biodiversity,
likelihood of extreme wildfires that can contaminate forest water supplies. Studies show or they provide benefits to society that are
that controlled burns do not degrade water quality compared with wildfires (Fernandes not easily quantified, such as opportunities
et al., 2013). for recreation, religious connection and
aesthetic enjoyment (Hallema, Robinne
and Bladon, 2018).

Unasylva 251, Vol. 70, 2019/1


60

Box 2
Longleaf pine restoration increases surface water delivery in the Altamaha
River basin in Georgia, United States of America
Longleaf pine (Pinus palustris) coverage in the southeastern Coastal Plain region of the United States of America declined in past centuries
from 372 000 km2 to 17 000 km2 due to agricultural conversion and replacement with loblolly pine (Pinus taeda) plantations. Natural longleaf
pine forest grows as savanna, with lower evapotranspiration, lower water demand and greater drought tolerance than dense loblolly pine for-
est. To assess the potential impacts of longleaf pine restoration on water, we simulated the 36 670 km2 Altamaha River basin for the period
1981–2010 using the Soil Water Assessment Tool. We compared water balances for the existing mixed land-use situation (34.3 percent ever-
green forest, 23.5 percent farmland, 22.1 percent deciduous forest, 11.6 percent wetland forest and 8.5 percent urban) with a scenario in which
all farmland was converted to loblolly pine (maximum seasonal leaf area index 5.0; Sampson et al., 2011) and another scenario in which all
farmland was converted into open longleaf pine savanna (leaf area index 2.0; Kao et al., 2012). The mixed land-use situation and the loblolly
pine and longleaf pine scenarios provided 486 mm, 430 mm (11.4 percent) and 498 mm (2.6 percent) of water yield, respectively, for 1 185 mm
average annual precipitation. Evapotranspiration was 671 mm (reference), 729 mm (8.6 percent) and 658 mm (2.0 percent), respectively. Given
declining annual precipitation and increased summer drought in the Southeast Region of the United States of America, a primary land man-
agement objective of longleaf pine restoration, combined with prescribed burning, would have a positive impact on surface water supplies.
© DENNIS HALLEMA, USDA FOREST SERVICE SOUTHERN RESEARCH STATION

© BRIAN BROWN, VANISHING SOUTH GEORGIA


Natural longleaf pine savanna in the southeast of the United Upstream forest restoration efforts have the potential to
States of America has an open canopy and does not consume as increase streamflow in the Altamaha River in Georgia,
much water as much denser loblolly pine forests United States of America

Under the right conditions, forests that 82 percent of the global population water quality due to upstream disturbances.
can supply high-quality drinking water uses water from upstream areas faced Overall, the ongoing decline in
with minimal treatment. A substantial with high levels of threat (Green et al., water quality is concerning, given accel-
part of the cost of water supply is gener- 2015). Remediation and purification erating trends in urbanization and water
ally associated with water purification efforts to safeguard water quality benefit demand (Sun, Hallema and Asbjornsen,
(Millennium Ecosystem Assessment, 2005); 75 percent of the population, but these ben- 2017), and it raises the question of
surface water supplies from undisturbed efits are unequally distributed: industrial how the cost of watershed protection and
forests that yield high-quality water countries reduce freshwater threats by aquifer recharge can be reduced (Muñoz-
usually have lower treatment costs 50–70 percent, while countries with Piña et al., 2008). In some cases, forest
compared with water from other sources lower gross domestic products reduce restoration could lead to an increase in
(García Chevesich et al., 2017). threats by less than 20 percent (Green et water supplies in the long term, even if
It’s easy to take clean water for granted al., 2015). it does not specifically target water ser-
when it is available in abundance. Nearly This disparity is linked not only to politi- vices (Box 2).
all forest watersheds are subject to some cal and economic factors but also to the
degree of human activity, however, degree of urbanization. Rapidly growing
and water scarcity and water impair- water-dependent urban centres are likely
ment are widespread. It is estimated to experience an increased risk of impaired

Unasylva 251, Vol. 70, 2019/1


61

WILDFIRE IMPACTS ON WATER


Box 3
SUPPLY SERVICES
Post-wildfire erosion in Chile and concerns for water supplies
Although wildfires have beneficial effects South-central Chile experienced major wildfires in 2017 that burned more than 5 000 km2.
on forest landscapes, the outcome can An unusually hot spring season combined with prolonged drought (Garreaud et al., 2017)
be very different for extreme wildfires triggered a series of fire storms. Approximately half of these occurred in radiata pine (Pinus
that consume forest stands – including radiata) plantations, and most were ignited by humans. In addition to the devastating effects
canopies – in their entirety. Wildfires tend on the human population and regional economy, there are serious concerns for biodiversity,
to increase storm runoff in the months after given that some burned areas are already on the International Union for Conservation of
a fire and boost the water yield from burned Nature’s Red List of ecosystems in critical danger of collapse (Alaniz, Galleguillos and
landscapes for several years (Kinoshita Perez-Quezada, 2016). The 2017 wildfires have increased erosion rates, even removing the
and Hogue, 2011; Kinoshita and Hogue, entire topsoil layer in some areas. This has led to the compaction of the now-exposed lower
2015; Hallema et al., 2017b; Hallema et al., soil layers due to the combined effect of relatively short forest rotation cycles (with as little as
2018). They also have profound impacts 20 years between harvests) and the higher impact force of raindrops on the now unvegetated
on the water purification functions of – and unprotected – soil surface (Soto et al., 2019). The phenomenon has reached a stage at
watersheds by changing the timescales which no more loose sediment is available for erosion, and the soil is effectively depleted.
and pathways of water movement through The concerning impact of wildfire in Chile shows the urgency of integrating water-related
landscapes and increasing the availability issues in sustainable forest management. It also demonstrates the need to further investigate
of readily transported material such as post-fire drainage issues and dissolved chemicals and suspended sediments that affect treatment
wildfire ash (Hallema et al., 2017a; Murphy processes for municipal water supplies (Odigie et al., 2016).
et al., 2018). Wildfire ash contains trace
metals, nutrients and organic material from
© MAURICIO GALLEGUILLOS, UNIVERSIDAD DE CHILE

branches, leaves and needles that can com-


promise water treatment for domestic uses.
Precipitation drives the transportation of
contaminants, ash and eroded soil down-
hill, resulting in pulses of increased stream
levels immediately following rainstorms
(Ice, Neary and Adams, 2004).
Combined with the loss of riparian veg-
etation and increased sediment loads in
streams, severe wildfires degrade aquatic
habitat and affect fisheries, which provide
important hydrological services and fulfil
vital economic roles in many parts of the
Intense fire storms in south-central Chile in 2017 caused a major loss of Pinus radiata
world. Locally, increased stream tempera- forest cover
tures and toxicity from ash, fire retardant
and polluted sediments are direct causes sediment reduces the capacity of reservoirs piping, pumps and filtration equipment,
of mortality among fish and other aquatic to store water, and adsorbed nutrients like and stabilizing streambanks and hillslopes
organisms (Dunham et al., 2007). phosphorus can promote algal growth – can cost millions of dollars (Box 4).
Degraded surface runoff can be conveyed (Smith et al., 2011). Studies in Australia
towards water intakes and water-storage and Chile have observed that fire-affected A HEALTHY FIRE REGIME FOR
reservoirs, often located at considerable water contains dissolved chemicals and SUSTAINABLE FORESTS AND
distances downstream from burned water- suspended sediments that affect treatment WATER SUPPLIES
sheds. For example, runoff from the 1996 processes for municipal water supplies Forests are resilient to and often benefit
Buffalo Creek Fire in Colorado, United and has the potential to affect human from fire, which promotes new growth
States of America, travelled more than health (White et al., 2006; Odigie et al., and species diversity and increases their
15 km from the burned area to a down- 2016) (Box 3). Measures to restore water natural ability to improve water quality by
stream reservoir (Moody and Martin, supply infrastructure after wildfire and soil filtration. Forests burned by extreme
2001). Floating debris clogs water intakes post-wildfire flooding – such as remov- wildfire ultimately recover the capacity to
and hydroelectric-generation equipment, ing sediment from reservoirs, repairing provide clean water, but the process can

Unasylva 251, Vol. 70, 2019/1


62

Box 4
Degraded post-wildfire water quality in urban water systems in California,
United States of America

The state of California is experiencing increasing fire risk due to warmer and drier conditions, yet urban development continues to encroach on
surrounding wildlands, exposing residents to growing primary and secondary fire hazards. The October 2017 North Bay wildfires in the San
Francisco Bay Area caused 46 fatalities and the loss of thousands of structures. These extreme fires, also known as the Northern California
Firestorm, constitute one of the state’s costliest disasters. One of the fires, the Tubbs Fire, damaged the drinking-water system, resulting in elevated
levels of benzene and other contaminants to the
extent that a local “do-not-drink/do-not-boil”
water-quality advisory was maintained until
one year after the fire. In Northern California,
drinking water in the city of Paradise became
contaminated with benzene after the 2018
© ZACK MONDRY, USDA FOREST SERVICE

Camp Fire, when burned plastics, soot and ash


leaked into the water system. It is estimated
that it could take two years and cost USD 300
million to restore the system’s water quality.
These examples highlight the detrimental
impacts of major wildfires on water quality
and demonstrate the need to protect drinking
water from future wildfires.
A severe wildfire in California destroyed much of the chaparral vegetation, leading to
erosion and increased sediment input in surface water

take many years (Robichaud et al., 2009). the resilience of forest water supplies (Robinne et al., 2016). As the timing,
Sustainable forest planning and manage- (Kinoshita et al., 2016; Hallema et al., magnitude and interaction of wildfires,
ment can mitigate the adverse impacts of 2019). droughts and insect infestations continue
extreme wildfires while helping maintain The future reliability of water supplies to change, additional alterations to forest
forest health and safeguarding forest water also depends on forest structure and veg- structure and function can be expected.
services (Postel and Thompson, 2005). A etative composition and their interactions More research is needed to better under-
healthy fire regime is the cornerstone of a with ecosystem processes (Thompson et stand the precursors of these unprecedented
sustainable forest and therefore a sustain- al., 2013). Increasing variability in air events to allow land managers to develop
able water supply (Figure 2). Promoting the temperature, precipitation, land use and and apply adaptive conservation practices
use of prescribed fire in watersheds can chemical deposition (nitrogen and sulphur) aimed at increasing hydrological resilience
reduce the likelihood of extreme wildfires is creating unprecedented combinations to forest disturbance.
and the consequent contamination of for- of ecosystem stress (McNulty, Boggs
est water supplies (Boisramé et al., 2017). and Sun, 2014), which can contribute to SAFEGUARDING FUTURE WATER
Given predictions that wildfires will changes in fire regimes and water cycles RESOURCES
increase in frequency, intensity and size in that are difficult to predict. In Cape Viewing the fire, water and society nexus
future climate regimes linked with increas- Province, South Africa, for example, the as a dynamic process helps in identify-
ing drought, scientists, policymakers and introduction of non-native acacias, euca- ing high-priority issues for scientists,
managers must coordinate their efforts in lypts and pines has increased fuel loadings, land managers and water providers. The
fire preparedness (warning systems), fire leading to increased fire risk (Kraaij et al., importance of this dynamic interaction is
impact planning and post-fire risk assess- 2018) and the possibility of post-fire water reflected in the International Association
ment to anticipate the potential post-fire quality effects. of Hydrological Sciences’ decadal
impacts on water. A good understanding Ultimately, increasing fire frequency and (2013–2022) research theme, Panta Rhei
of fire trends, impacts and environmental severity affect the quality and quantity (“everything flows”). Forest disturbances
interactions is essential for maintaining of forest water resources at broad scales accumulate downstream, and therefore the

Unasylva 251, Vol. 70, 2019/1


63

MARKET NON-MARKET
EROSION
DRINKING WATER AQUATIC ECOSYSTEM
FLOODING QUALITY FUEL TREATMENTS
HYDROPOWER (PRESCRIBED BURNING)
CONTAMINATION WITH
METALS, NUTRIENTS, ASH POWER PLANT BIODIVERSITY, REFORESTATION
COOLING RECREATION
RESERVOIR POLLUTION ADAPTIVE
IRRIGATION, RELIGIOUS, CONSERVATION
AQUACULTURE AESTHETIC
FLOOD
MITIGATION

2
Wildfires can have a severe impact on water
services, but much of this impact may be
mitigated by fuel treatment and other forest
solutions for water (Robinne et al., 2018) consumed by fire. The challenge is that
management practices and to integrate a more fundamental every fire has unique circumstances, and
understanding of interactions between ground data are scarce.
future of water resources is the inevitable wildfires, reforestation/afforestation, and The trend of increasing urbanization will
sum of natural and human impacts and the supply of and demand for hydrological lead to more deforestation and increase
their interactions and feedbacks. services (Box 5). pressure on forest hydrologic services.
The quality of water-supply predictions Expanding the area of study from the Two-thirds of the global population is
depends in large part on the quality of local to regional scale has major implica- expected to reside in urban areas by 2050,
data and models. The wealth of satellite tions for the number of interactions that with most growth concentrated in Africa,
data on wildfires, climate and forest inven- must be taken into account. To quantify Asia, Latin America and the Caribbean
tory collected in recent years has enabled fire risk to water security, for example, (UN-Habitat, 2018). The land area covered
the building of predictive models of fire it is necessary to identify “at risk” for- by cities is predicted to triple, and more
impacts on water. Few datasets exist, ests where active management is needed people are expected to move into the transi-
however, on post-fire water quality, and to safeguard water supplies and public tion zone between forests and urban areas.
predictive models rely on ground data for health. This requires the involvement of The take-away is that wildland fire
validation, which is often a challenge in forest managers, hydrologists, wildfire impacts on water supply and water
developing countries. Although higher scientists, public-health specialists and quality will continue to extend well
spectral and temporal data resolutions are the public. There is also a need to quantify beyond forest boundaries and to
a welcome development, scientists need water contamination coming from burnt directly affect the forest hydrologic
better training in the use of these data to anthropogenic sources such as plastics, services of people living downstream.
predict the effectiveness of nature-based gases and fabrics when builtup areas are Ultimately, a better understanding of

Unasylva 251, Vol. 70, 2019/1


64

Box 5
China’s “Grain-for-Green” programme: improving water quality through
afforestation and forest restoration

Satellite imagery shows that China is becoming greener following years of afforestation and forest protection efforts. The aim of the
Conversion of Cropland to Forest Programme (CCFP), or “Grain-for-Green”, the world’s largest payment scheme for ecosystem services, is
to combat soil erosion and improve the rural environment. Afforestation (planting trees where no forest existed previously) is one of its core
activities, financed through a public payment scheme that involves millions of rural households (Lü et al., 2012). Sediment monitoring in
the Yangtze River and elsewhere shows evidence of reduced sediment loads following the start of the CCFP in 1999 and the Natural Forest
Protection Programme in 1998, with a positive effect on drinking-water quality (Zhou et al., 2017; Mo, 2007). There are concerns, however,
that afforestation with non-native tree species uses too much water and causes soil desiccation (Deng et al., 2016), potentially leading to lower
water levels in, for example, the Yellow River, with severe consequences for downstream water supply. Additionally, forest planning in China
has rarely considered prescribed burning as a management tool and instead favours fire suppression. There is a strong need to monitor and
predict potential fire impacts on water services to ensure the cost-effectiveness of forest restoration efforts (Cao et al., 2011).

© STACIE WOLNY, STANFORD UNIVERSITY

Forest restoration in southern China’s Pearl River Basin has reduced erosion, leading to
better water quality in rivers

regional fire impacts and interactions is ACKNOWLEDGEMENTS (ORISE) through an interagency agreement
needed for a breakthrough in the devel- Dennis W. Hallema received support from between the United States Department
opment of cost-effective strategies for the Forest Service Research Participation of Energy (DOE) and the USDA Forest
managing fire and water. Program administered by the Oak Ridge Service. ORISE is managed by Oak
Institute for Science and Education Ridge Associated Universities (ORAU)

Unasylva 251, Vol. 70, 2019/1


65

under DOE contract number DE-AC05- Deng, L., Yan, W., Zhang, Y. & Shangguan, specific connectivity. Hydrological
06OR23100. Funding was provided by the Z. 2016. Severe depletion of soil moisture Processes, 31(14): 2582–2598.
South Atlantic Landscape Conservation following land-use changes for ecological Hallema, D.W., Sun, G., Caldwell, P.V.,
Service through an interagency agreement restoration: evidence from northern China. Norman, S.P., Cohen, E.C., Liu, Y.Q.,
between the USDA Forest Service and the Forest Ecology and Management, 366: 1–10. Ward, E. J. & McNulty, S.G. 2017b.
United States Fish and Wildlife Service. Dunham, J.B., Rosenberger, A.E., Luce, C.H. Assessment of wildland fire impacts on
Any opinions, findings, conclusions & Rieman, B.E. 2007. Influences of wildfire watershed annual water yield: Analytical
or recommendations expressed in this and channel reorganization on spatial and framework and case studies in the United
article are those of the authors and do not temporal variation in stream temperature States. Ecohydrology, 10(2): 20.
necessarily reflect the policies and views and the distribution of fish and amphibians. Hallema, D.W., Sun, G., Caldwell, P.V.,
of the Government of the United States Ecosystems, 10(2): 335–346. Norman, S.P., Cohen, E.C., Liu, Y.,
of America. Any use of trade, firm or Ebel, B.A. & Moody, J.A. 2017. Synthesis Bladon, K.D. & McNulty, S.G. 2018.
product names is for descriptive purposes of s oi l- hyd r a u l ic p r o p e r t ie s a n d Burned forests impact water supplies. Nature
only and does not imply endorsement by infiltration timescales in wildfire-affected Communications, 9(1): 1307.
the Government of the United States of soils. Hydrological Processes, 31(2): 324– Hallema, D.W., Sun, G., Caldwell, P.V.,
America. u 340. Robinne, F.-N., Bladon, K.D., Norman,
Fernandes, P.M., Davies, G.M., Ascoli, D., S.P., Liu, Y., Cohen, E.C. & McNulty, S.G.
Fernández, C., Moreira, F., Rigolot, E., 2019. Wildland fire impacts on water yield
Stoof, C.R., Vega, J.A. & Molina, D. 2013. across the contiguous United States. Gen.
Prescribed burning in southern Europe: Tech. Rep. SRS-238. Asheville, USA, US
developing fire management in a dynamic Department of Agriculture Forest Service
References landscape. Frontiers in Ecology and the Southern Research Station.
Environment, 11(s1): e4–e14. Ice, G.G., Neary, D.G. & Adams, P.W. 2004.
Alaniz, A.J., Galleguillos, M. & Perez- García Chevesich, P., Neary, D.G., Scott, Effects of wildfire on soils and watershed
Quezada, J.F. 2016. Assessment of D.F., Benyon, R.G. & Reyna, T. 2017. processes. Journal of Forestry, 102(6): 16–20.
quality of input data used to classify Forest management and the impact on Kao, R.H., Gibson, C.M., Gallery, R.E.,
ecosystems according to the IUCN Red water resources: a review of 13 countries. Meier, C.L., Barnett, D.T., Docherty,
List methodology: the case of the central UNESCO Publishing. K.M., et al. 2012. NEON terrestrial field
Chile hotspot. Biological Conservation, Garreaud, R.D., Alvarez-Garreton, C., observations: designing continental-
204: 378–385. Barichivich, J., Boisier, J.P., Christie, D., scale, standardized sampling. Ecosphere,
Balch, J.K., Bradley, B.A., Abatzoglou, Galleguillos, M., LeQuesne, C., McPhee, 3(12): 1–17.
J.T., Nagy, R.C., Fusco, E.J. & Mahood, J. & Zambrano-Bigiarini, M. 2017. The Kinoshita, A.M., Chin, A., Simon, G.L.,
A.J. 2017. Human-started wildfires expand 2010–2015 megadrought in central Chile: Briles, C., Hogue, T.S., O’Dowd, A.P.,
the fire niche across the United States. impacts on regional hydroclimate and Gerlak, A.K. & Albornoz, A.U. 2016.
PNAS, 114(11): 2946–2951. doi:10.1073/ vegetation. Hydrology & Earth System Wildfire, water, and society: toward
pnas.1617394114 Sciences, 21(12): 6307–6327. integrative research in the “Anthropocene”.
Boisramé, G., Thompson, S., Collins, B. & Green, P.A., Vörösmarty, C.J., Harrison, I., Anthropocene, 16: 16–27.
Stephens, S. 2017. Managed wildfire effects Farrell, T., Sáenz, L. & Fekete, B.M. 2015. Kinoshita, A.M. & Hogue, T.S. 2011. Spatial
on forest resilience and water in the Sierra Freshwater ecosystem services supporting and temporal controls on post-fire hydrologic
Nevada. Ecosystems, 20(4): 717–732. humans: pivoting from water crisis to water recovery in Southern California watersheds.
Cao, S., Sun, G., Zhang, Z., Chen, L., Feng, solutions. Global Environmental Change, Catena, 87(2): 240–252.
Q., Fu, B., McNulty, S.G., Shankman, 34: 108–118. Kinoshita, A.M. & Hogue, T.S. 2015. Increased
D., Tang, J., Wang, Y. & Wei, X. 2011. Hallema, D.W., Robinne, F.-N. & Bladon, dry season water yield in burned watersheds
Greening China naturally. Ambio, 40(7): K.D. 2018. Reframing the challenge of in Southern California. Environmental
828–831. global wildfire threats to water supplies. Research Letters, 10(1): 014003.
Carvalho-Santos, C., Honrado, J. P. & Earth’s Future, 6(6): 772–776. Kraaij, T., Baard, J.A., Arndt, J., Vhengani,
Hein, L. 2014. Hydrological services and Hallema, D.W., Sun, G., Bladon, K.D., L. & van Wilgen, B.W. 2018. An assessment
the role of forests: conceptualization and Norman, S.P., Caldwell, P.V., Liu, Y. & of climate, weather, and fuel factors
indicator-based analysis with an illustration McNulty, S.G. 2017a. Regional patterns influencing a large, destructive wildfire
at a regional scale. Ecological Complexity, of post-wildfire streamflow in the western in the Knysna region, South Africa. Fire
20: 69–80. United States: the importance of scale- Ecology, 14(4).

Unasylva 251, Vol. 70, 2019/1


66

Lü, Y., Fu, B., Feng, X., Zeng, Y., Liu, Murphy, S.F., McCleskey, R.B., Martin, D.A., Smith, H.G., Sheridan, G.J., Lane, P.N.J.,
Y., Chang, R., Sun, G. & Wu, B. 2012. Writer, J.H. & Ebel, B.A. 2018. Fire, flood, Nyman, P. & Haydon, S. 2011. Wildfire
A policy-driven large scale ecological and drought: extreme climate events alter effects on water quality in forest catchments:
restoration: quantifying ecosystem services flow paths and stream chemistry. Journal a review with implications for water supply.
changes in the Loess Plateau of China. PloS of Geophysical Research: Biogeosciences, Journal of Hydrology, 396 (1–2): 170–192.
ONE, 7(2): e31782. 123. https://doi.org/10.1029/2017JG004349 Soto, L., Galleguillos, M., Seguel, O.,
Martin, D.A. 2016. At the nexus of fire, water Odigie, K.O., Khanis, E., Hibdon, S.A., Sotomayor, B. & Lara, A. 2019. Assessment
and society. Philosophical Transactions of Jana, P., Araneda, A., Urrutia, R. & of soil physical properties’ statuses under
the Royal Society B: Biological Sciences, Flegal, A.R. 2016. Remobilization of trace different land covers within a landscape
371(1696): 20150172. elements by forest fire in Patagonia, Chile. dominated by exotic industrial tree
McDonald, R.I., Weber, K.F., Padowski, J., Regional Environmental Change, 16(4): plantations in south-central Chile. Journal of
Boucher, T. & Shemie, D. 2016. Estimating 1089–1096. Soil and Water Conservation, 74(1): 12–23.
watershed degradation over the last century Paton, D., Buergelt, P.T., Tedim, F. & Sun, G., Hallema, D.W. & Asbjornsen,
and its impact on water-treatment costs for McCaffrey, S. 2015. Wildfires: international H. 2017. Ecohydrological processes and
the world’s large cities. Proceedings of the perspectives on their social–ecological ecosystem services in the Anthropocene: a
National Academy of Sciences, 113(32): implications. In: D. Paton, P.T. Buergelt, review. Ecological Processes, 6: 35.
9117–9122. F. Tedim & S. McCaffrey, eds. Wildfire Sun, G., Hallema, D.W. & Asbjornsen, H.
McNulty, S.G., Boggs, J.L. & Sun, G. hazards, risks and disasters, pp. 1–14. 2018. Preface for the article collection
2014. The rise of the mediocre forest: why London, Elsevier. “Ecohydrological processes and ecosystem
chronically stressed trees may better survive Pausas, J.G. & Keeley, J.E. 2019. Wildfires services”. Ecological Processes, 7: 8.
extreme episodic climate variability. New as an ecosystem service. Frontiers in Thompson, M.P., Marcot, B.G., Thompson,
Forests, 45(3): 403–415. Ecology and the Environment, 17(5): 289– F.R., McNulty, S., Fisher, L.A., Runge,
Melchiorre, A. & Boschetti, L. 2018. Global 295. M.C., Cleaves, D. & Tomosy, M. 2013. The
analysis of burned area persistence time with Postel, S.L. & Thompson Jr, B.H. 2005. science of decisionmaking: applications for
MODIS data. Remote Sensing, 10(5): 750. Watershed protection: capturing the benefits sustainable forest and grassland management
Millennium Ecosystem Assessment. of nature’s water supply services. Natural in the National Forest System. Gen. Tech.
2005. Ecosystems and human well- Resources Forum, 29(2): 98–108. Rep. WO-GTR-88. Washington, DC, US
being: synthesis. Washington, DC, Robichaud, P.R., Lewis, S.A., Brown, R.E. Department of Agriculture Forest Service.
Island Press. & Ashmun, L.E. 2009. Emergency post-fire UN-Habitat (United Nations Human
Mo, Z. 2007. Facilitating reforestation rehabilitation treatment effects on burned Settlements Programme). 2018. Working
for Guangxi watershed management area ecology and long-term restoration. Fire for a Better Urban Future Annual Progress
in Pearl River Basin Project. FAO Ecology, 5(1): 115–128. Report 2018 (available at https://unhabitat.
Advisory Committee on Paper and Wood Robinne, F.-N., Bladon, K.D., Miller, C., org/wp-content/uploads/2019/05/HSP-
Products, 48th Session, Shanghai, China, Parisien, M.-A., Mathieu, J. & Flannigan, H A-1-I N F 2-U N-H A BI TAT-A n nua l-
6 June 2007. M.D. 2018. A spatial evaluation of global Report-2018.pdf).
Moody, J.A. & Martin, D.A. 2001. Initial wildfire-water risks to human and natural White, I., Wade, A., Worthy, M., Mueller,
hydrologic and geomorphic response systems. Science of the Total Environment, N., Daniell, T. & Wasson, R. 2006. The
following a wildfire in the Colorado Front 610 – 611: 1193–1206. DOI:10.1016/j. vulnerability of water supply catchments
Range. Earth Surface Processes and scitotenv.2017.08.112 to bushfires: impacts of the January 2003
Landforms, 26(10): 1049–1070. Robinne, F.-N., Miller, C., Parisien, M.-A., wildfires on the Australian Capital Territory.
Moritz, M.A., Batllori, E., Bradstock, R.A., Emelko, M.B., Bladon, K.D., Silins, U. Australasian Journal of Water Resources,
Gill, A.M., Handmer, J., Hessburg, P.F., & Flannigan, M. 2016. A global index for 10(2): 179–194.
Leonard, J., McCaffrey, S., Odion, D.C., mapping the exposure of water resources to Zhou, Y., Ma, J., Zhang, Y., Qin, B.,
Schoennagel, T. & Syphard, A.D. 2014. wildfire. Forests, 7(1): 22. Jeppesen, E., Shi, K., Brookes, J.D.,
Learning to coexist with wildfire. Nature, Sampson, D.A., Amatya, D.M., Lawson, Spencer, R.G.M., Zhu, G. & Gao, G.
515(7525): 58. C. B. & Skaggs, R.W. 2011. L eaf 2017. Improving water quality in China:
Muñoz-Piña, C., Guevara, A., Torres, area index (LAI) of loblolly pine and environmental investment pays dividends.
J.M. & Braña, J. 2008. Paying for emergent vegetation following a harvest. Water Research, 118: 152–159. u
the hydrological services of Mexico’s Transactions of the American Society of
forests: analysis, negotiations and results. Agricultural and Biological Engineers,
Ecological Economics, 65(4): 725–736. 54(6): 2057–2066.

Unasylva 251, Vol. 70, 2019/1


67

Nature-based solutions for water-related disasters


L. Spurrier, A. Van Breda, S. Martin, R. Bartlett and K. Newman

© WWF-MADAGASCAR

M
Mangroves are crucial resources angrove forests are among the building materials and local subsistence
for many of the world’s most world’s most valuable coastal use (Nagelkerken et al., 2008).
vulnerable people, but more ecosystems, providing local Serving as a transition between marine
research and coordination is communities with numerous services and and terrestrial environments, mangroves
needed to inform decisions on their benefits. They function as nursery habitats can provide vulnerable people with protec-
role in disaster risk reduction. for many animals, such as crab, prawn and tion against the impacts of climate change
fish species, support local food webs, and (Munang et al., 2013) by attenuating wave
Lauren Spurrier is Vice President, Oceans
Conservation, World Wildlife Fund (WWF), create linkages with other ecosystems for energy and storm surges, buffering ris-
Washington, DC, United States of America nutrient cycling and migratory pathways ing sea levels, stabilizing shorelines from
Anita Van Breda is Senior Director, (Nagelkerken et al., 2008). Mangroves erosion, and contributing to general flood
Environment and Disaster Management, WWF,
Washington, DC, United States of America also support a vast array of culturally control (Vo et al., 2012). Mangroves pro-
Shaun Martin is Senior Director, Climate and environmentally important species, vide even more ecosystem services when
Change Adaptation and Resilience, WWF, including shorebirds, crocodiles, mana- coupled with coral reefs and seagrass beds.
Washington, DC, United States of America
Ryan Bartlett is Director, Climate Risk and tees, and even tigers in the Sundarbans Despite the many benefits, however, the
Resilience, WWF, Washington, DC, United States (Danda et al., 2017). Millions of people global extent of mangroves has declined
of America living along coasts benefit from mangroves
Kate Newman is Vice President, Forest and
Freshwater Public Sector Initiative, WWF, for aquaculture, agriculture, forestry, pro- Above: Community mangrove
Washington, DC, United States of America tection against shoreline erosion, woodfuel, restoration in Madagascar

Unasylva 251, Vol. 70, 2019/1


68

by 67 percent in the last century. Roughly DRR in a changing climate. It recom- mangroves should feature in overall disas-
1 percent of mangrove cover is now being mends further research, collaboration ter risk management strategies.
lost per year (FAO, 2007), driven by coastal and coordination among the humanitar- Traditional built infrastructure options
development, aquaculture, resource use ian, land-use planning, conservation, for coastal defence are proving inadequate
and, in some instances, climate change. climate-change adaptation, development, – and at times counterproductive – in deliv-
Mangrove loss can increase the vulner- and disaster management sectors to enable ering risk-mitigation outcomes to life and
ability of coastal communities and the risks better-informed decisions on the use of property. Hard defences such as sea walls,
to which they are exposed (Blankespoor, mangroves for DRR. levees and bulkheads can provide a general
Dasgupta and Lange, 2016). Continued sense of security because they are familiar,
declines in mangrove area could lead to COASTAL PROTECTION FROM A well understood and often constructed
dramatic losses of biodiversity, increased COMBINATION OF NATURAL AND following codified local regulations. For
salt intrusion in coastal zones, and the BUILT INFRASTRUCTURE these reasons and others, hard infrastruc-
siltation of coral reefs, ports and shipping Integrating mangroves in DRR is not an ture is often preferred over nature-based
lanes, with consequent losses of income entirely new concept. Coastal managers and solutions like mangroves. Nevertheless,
and livelihood options (FAO, 2007). scientists have long recognized the value of there are many disincentives for relying
Mangroves are essential for reducing the mangroves and related coastal ecosystems, solely on built infrastructure. For example,
vulnerability of many coastal communi- such as coral reefs and seagrass beds, in built infrastructure has high construction,
ties to the impacts of climate change and mitigating the impacts of coastal hazards, operational and maintenance costs, and it
increasingly intense and frequent extreme including storm surges and flooding from is strongest immediately after construc-
weather events. Yet climate change itself cyclones and, to some extent, tsunamis. As tion and then weakens with age. Moreover,
presents a significant threat to mangroves coastal “bioshields”, mangrove forests can hard infrastructure is built using specific
that could undermine their value in reduc- attenuate wave energy and reduce vulner- parameters, and it might be difficult to
ing this vulnerability (Algoni, 2015). To ability to storm surge inundation, although adapt it to rising sea levels or other changed
maximize opportunities for disaster risk their effectiveness depends on a range of conditions. It can also cause coastal habi-
reduction (DRR), conservationists and site-specific factors (Box 1). Variations in tat loss and have negative impacts on the
disaster risk managers must pay careful coastal characteristics and the influence
attention to the threats that are driving the of humans on mangrove systems affect
rapid loss of mangroves globally and their the relative value of mangroves as coastal Mangroves in Sundarban Forest on the edge
capacity to survive in a changing climate. defence mechanisms. Thus, it is impor- of the Bay of Bengal, Khulna Province, south
coast of Bangladesh. These mangroves
Disaster risk managers must also con- tant to carefully review sites to determine are battered by the sea, but they play an
sider the current and future viability of the extent to which existing or restored important role in protecting the coast from
the protective services of mangroves. For storms and erosion
example, the implications of the combined
impacts of sea-level rise, changing salin-
ity, extreme weather events, economic
development (e.g. aquaculture and fisher-
ies) and infrastructure development (e.g.
roads, dams and urbanization) should be
understood to best determine how man-
groves might contribute to risk reduction
for people in long-term DRR planning. At
the same time, conservationists must take
urgent action to reduce threats to existing
mangroves and to enable mangrove tree
© NATUREPL.COM/TIM LAMAN/WWF

species to migrate inland and to new areas


as sea levels rise. Policymakers and land
managers must understand the interaction
of these factors at a landscape scale.
This article presents the factors that
should be considered when evaluating
the value of mangrove ecosystems for

Unasylva 251, Vol. 70, 2019/1


69

Box 1
Planning considerations
A comprehensive understanding of mangrove ecosystems is crucial for an agency’s (governmental or community-managed) capacity to plan
the role of mangroves in DRR strategies. The duration and extent to which mangroves can provide protection depends on many factors,
including – but not limited to – the following:
• the characteristics of mangroves and the surrounding environment;
• the physical and geological conditions of a site, such as forest floor shape, shoreline configuration and bathymetry;
• the size of the ecosystem, vegetation density and stiffness (contributes to an understanding of frictional resistance), and the height of the
mangrove forest (Sutton-Grier, Wowk and Bamford, 2015); and
• the hydrologic functioning of the landscape (Radabaugh et al., 2019).

Characteristics of threats:
• The spectral characteristics of the incident waves and the tidal stage at which a wave enters a forest (Blankespoor, Dasgupta and Lange, 2016)
• The height of the storm wave and windspeed (Spalding et al., 2014)
• The number and duration of extreme events (Spalding et al., 2014)
• Mangrove resilience, recovery and regeneration following an extreme event
• The type of hazard to which a community is exposed (e.g. storm, tsunami, erosion, sea-level rise) (Spalding et al., 2014)
• The distance of human communities and infrastructure from the shoreline
• The human response to the event (e.g. cutting poles for rebuilding, the placement of hard structures or barriers within a mangrove eco­
system, and the rebuilding of roads adjacent to the ecosystem).

ecosystem services provided by nearby defence capacity and deliver a wide range impacts in the Lower Florida Keys and
coastal ecosystems (Sutton-Grier, Wowk of co-benefits that contribute to the overall Ten Thousand Islands. It made several key
and Bamford, 2015). economic, social and ecological resilience findings, including:
Some of the limitations of built (“grey”) of coastal systems. • There was extensive canopy damage
and natural (“green”) infrastructure can from high winds; the canopy cover
be addressed by using a hybrid approach DISASTER IMPACTS ON increased from 40 percent to 60 per-
combining grey and green coastal defence MANGROVES AND POST-EVENT cent within 2–4 months after the hurri-
options in shoreline management, par- RECOVERY cane but recovery plateaued thereafter.
ticularly in cyclone and tsunami-prone Effective DRR planning should consider • Deposits of mud and debris (which
areas. According to a growing body of the impacts of extreme events on man- could include non-organic debris,
research and experimentation globally, groves. The rate and extent of recovery can household items and furniture) from
the combination of natural and built infra- vary widely depending on factors such as storms hamper regrowth by smoth-
structure can both increase resilience and species type, sediment availability, tem- ering roots and soil and decreasing
decrease costs. For example, mangrove perature, precipitation, storminess and oxygen exchange. Trees that initially
projects in Viet Nam can be 3–5 times sea-level rise (Ward et al., 2016). Some survive a storm may die due to this
cheaper (depending on water depth) than a extreme events may completely destroy smothering.
breakwater for the same level of protection an area of mangroves, transforming it • A lack of water, or excess water, can
(Narayan et al., 2016). Hybrid approaches into mudflats (Smith et al., 2009). Meta- kill mangroves.
can enhance natural systems and their analysis of disaster research (Mukherjee • Forests with appropriate elevation,
benefits. Newly restored mangroves, for et al., 2010) suggests that, despite many hydrology and a source of propagules
example, can be weak while tree roots post-disaster assessments, few include a should recover naturally.
take hold, and younger trees have a bet- comprehensive analysis of disaster impacts Understanding the impacts of extreme
ter chance of surviving major storms and on mangroves and related ecosystems or events on coastal ecosystems like man-
other stressors if protected by permeable combine natural, social, economic and risk groves, as in the above example after
(temporary) engineered structures as they management analysis. As a result, there Hurricane Irma, and how they recover from
mature (Sutton-Grier, Wowk and Bamford, is no clear understanding of how coastal such events, can provide a basis for deter-
2015). By ensuring that mangroves are systems collectively provide DRR services. mining future mangrove risk-reduction
restored and maintained, coastal managers A recent report (Radabaugh et al., potential. The post-disaster monitoring
and decision-makers can increase coastal 2019) documented post-Hurricane Irma and analysis of mangrove ecosystems can

Unasylva 251, Vol. 70, 2019/1


70
© ADAM OSWELL/WWF-GREATER MEKONG

A mangrove tree in the Mekong delta,


Viet Nam, with its roots extending 5m up
the trunk

also inform decisions on the suitability INCREASED ATMOSPHERIC CO


²
and value of assisting regeneration pro-
cesses and provide insights useful for the
development of future DRR objectives and
activities. INCREASE IN TEMPERATURE
The case of Cyclone Jokwe in
Mozambique in 2008 illustrates how
other post-disaster challenges are often
not considered. At the request of CARE, INCREASED SEA-LEVEL OCEAN ALTERED
PRECIPITATION
WWF conducted a rapid environmental STORMINESS RISE CURRENTS
REGIME
assessment in the immediate aftermath
of the cyclone to examine the potential for
CARE’s reconstruction strategy to include
environmentally responsible approaches

PLANT EROSION SEDIMENT MANGROVE SALINITY


PRODUCTIVITY SUPPLY DROWNING
1
Main factors and pathways affecting
mangrove systems under climate change Source: Ward et al. (2016), used here under a CC BY 4.0 licence.

Unasylva 251, Vol. 70, 2019/1


71

© ANTONIO BUSIELLO/WWF-US
The development of a tourism project is
destroying mangroves at Harvest Cayes,
levels; 2) increasing atmospheric carbon 2016) estimated the loss of protective
Placencia, Belize, Central America
dioxide; 3) warmer air and water tempera- services from mangroves in 46 countries
tures; 4) changing ocean currents; and 5) under a future scenario of a 1 m rise in sea
and to support joint CARE–WWF natu- the increasing variability and intensity of level and a 10 percent increase in storm
ral resource management activities. The rainfall (see Figure 1 on p. 68). Of these, intensification. It found that, assuming
subsequent report showed that extraction sea-level rise is the most significant chal- no loss of mangroves to human action
pressure on mangrove systems can increase lenge because it increases soil salinity and or sea-level rise, coastal flooding would
following a disaster: communities began thus drives higher rates of seedling mortal- increase by only 2 percent globally. Results
rebuilding their homes immediately using ity (Ward et al., 2016). How and whether were dramatically different and varied
mangrove timber, driving up the rate of mangroves survive in a future of increasing significantly by country, however, when the
mangrove consumption more than 14-fold climatic change and sea-level rise will be loss of mangroves due to the 1 m rise was
compared with non-emergency times, determined by four factors: 1) whether factored in. Indonesia, for example, would
potentially reducing the protective role of they can migrate inland (depending on the lose about 17 percent of its mangroves to
the mangrove ecosystem. DRR managers topography or infrastructure in their way); sea-level rise, thus causing the loss of
should factor such issues into planning 2) tidal range location and geomorphic coastal protection in those areas. Mexico,
and management. setting (related to the type of mangrove on the other hand, was estimated to lose
community); 3) continued supplies of sedi- all its existing mangroves, suggesting that
MANGROVE VULNERABILITIES DUE ment; and 4) whether mangrove migration investments in mangroves for coastal pro-
TO CLIMATE CHANGE inland can outpace the rate of sea-level rise tection in that country might not deliver
The conservation and disaster risk man- (Blankespoor, Dasgupta and Lange, 2016; expected benefits in the mid to long term.
agement communities often disregard the Ward et al., 2016). The study by Blankespoor, Dasgupta
vulnerability of mangroves to climate Given the vulnerability of mangroves to and Lange (2016) has several limitations,
change in planning substantial investments sea-level rise in some locations, appropri- and it could have over- or underestimated
in mangrove protection and restoration ate site selection is crucial for ensuring the the future protective value of mangrove
(Sutton-Grier, Wowk and Bamford, 2015). continued delivery of coastal protection ecosystems. Nevertheless, its findings
Mangroves are directly affected by climate services from mangroves. A World Bank demonstrate an often-overlooked point:
change through five vectors: 1) rising sea study (Blankespoor, Dasgupta and Lange, although mangroves will be crucial

Unasylva 251, Vol. 70, 2019/1


72

Mozambique
Mangrove cover reduced byBox 20,8932ha between 1996 and 2016 in Mozambique representing 6.5% of 1996 mangrove cover (Figure
Ecosystem
5). There are 18linkages: mangroves
rivers with a Vol/Len>100 and cover
with mangrove riverswithin the 10x AOI exceeding 10 ha in either 1996 or 2016. Eleven of
these rivers are free-flowing, two have good connectivity, and 5 are not free-flowing. Two of the 18 rivers are river order 3, eleven
WWF identified linkages between rivers and 2 ha, -3.8%,
are order 4, and five are river order 5. The greatest loss of mangroves near river-ocean outlets occurred by Zambezi (-1304
river #13 in Figure 5). The greatest gain in mangrove area occurred by Pungue (river #14 in Figure 5, a 201 ha gain in area,
mangroves at the national and global scale in a
representing 9.4% of 1996 mangrove area). River–mangrove connectivity in Mozambique
recent project (Maynard et al., 2019) by combining
two novel, state-of-the-art global datasets on rivers
(Grill et al., 2019) and mangrove extent in 1996–
2016 (Bunting et al., 2018). Rivers were catego-
rized as “free-flowing” (very few human impacts),
“good connectivity” (slight human impacts), or
“not free-flowing” (moderate to severe human
impacts). The project indicated that sediment trap-
ping by dams was a major driver of mangrove loss
adjacent to rivers, reducing the DRR potential of
these forests. Further analysis is ongoing.1 Here
we present the results from Mozambique.
Mozambique experienced a national decline
in mangrove cover from 319 445 ha in 1996 to
298 552 ha in 2016, equivalent to an overall loss
of 7 percent. Mangroves adjacent to rivers were
lost at a lower rate (3 percent), from 55 853 ha in
1996 to 54 389 ha in 2016. Of 18 selected rivers
in Mozambique (Figure 2), 11 were categorized
as “free-flowing” and two as having “good con-
nectivity”. Five rivers, however, were categorized
as “not free-flowing”, with significant human
impacts. All five of these rivers are in southern
Mozambique and have experienced substantial
losses of mangrove area: for example, the Zambezi
(River 13 in Figure 2) lost 1 304 ha of mangroves
between 1996 and 2016. With increasing energy
demand, Mozambique is looking to expand river
hydroelectric power generation, with four new
dams planned nationally for the Zambezi (Zarfl
et al., 2015). Moreover, an additional 12 dams
are planned upstream, where the Zambezi and
Figure 5.Maynard
Source: Country et map summary for Mozambique; mangrove cover change between 1996 and 2016
al. (2019).
its tributaries flow through Malawi, the United
near river-ocean outlets.
Republic of Tanzania, Zambia and Zimbabwe Notes: All rivers with a cross-sectional area greater than 100 m² at the river–ocean outlet
(this can be visualized as a river 50 m wide and 2 m deep) are numbered. River order reflects 19
(Zarfl et al., 2015). These dams are likely to further how a river fits within a river network, with the longest continuous river from source to ocean
exacerbate mangrove loss in the Zambezi delta. identified as order 1. Tributaries off this backbone are labelled as order 2 along their longest
length from source to backbone; further tributaries from these order 2 river segments are
labelled order 3, 4, etc., to a maximum of 7 river orders.

in many areas for protecting coastal MANGROVE VULNERABILITIES DUE easily be undermined by external threats
assets and people from increasingly TO UPSTREAM DEVELOPMENT that occur beyond their “boundaries”. This
extreme storms and flooding, advocates Mangrove ecosystems are highly connected is especially important at river–ocean
should be careful not to oversell the to adjacent landscapes and seascapes, boundaries, where processes occurring
benefits, especially in areas with low which affect their health and integrity far upstream in terrestrial watersheds can
potential for inland mangrove migration (Ervin et al., 2010). Attempts to conserve have large impacts on coastal ecosystems.
(such as in Mexico). mangrove forests locally, therefore, can For example, 31 percent of sediment that

Unasylva 251, Vol. 70, 2019/1


73

should be carried into Asian deltas and ecosystems but should take a carefully • As mangroves migrate due to a chang-
which would help replenish coastlines is designed, hybrid approach that includes ing climate, they may no longer pro-
prevented by dams (Syvitski et al., 2005), appropriate built infrastructure, as well vide the same protective functions to
with major rivers in Pakistan, Thailand as early-warning systems and community communities relying on them.
and Viet Nam experiencing 75–95 percent education (Blankespoor, Dasgupta and • Adaptive frameworks and decision-
declines in coastal sediment deposition Lange, 2016). support tools that enable managers to
(Gupta, Kao and Dai, 2012). Such upstream It is especially important to factor integrate and continuously update pro-
sediment-trapping leads to increased in climate change (including sea-level jections of climate-change risk, land
coastal erosion and causes mangrove for- rise) because it may cause mangroves to use and human population growth can
est loss, with examples observed on the migrate and thereby reduce their protec- increase the effectiveness of natural
Mexican Pacific coast (Ezcurra et al., 2019) tive function in a particular area. Adaptive infrastructure, including mangroves.
and in the Mekong delta (Li et al., 2017) frameworks and decision-support tools • Although the value of mangroves in
and Mozambique (Box 2). Yet coastal can increase the effectiveness of natural providing DRR services are well docu-
DRR efforts and upstream river man- infrastructure, including mangroves, by mented, they are rarely considered in
agement efforts are routinely disjointed, enabling managers to integrate and con- planning and development. Neverthe-
with mangrove forest jurisdiction often tinually update the risks posed by climate less, in some circumstances, assertions
slipping through the cracks between ter- change as well as changes in land use and that mangroves reduce disaster risk
restrially focused government forestry and human populations (Powell et al., 2019). are overly simplistic, do not present
environment agencies and marine- and In summary: the full suite of conditions and issues
fisheries-focused departments. Without • There is a growing body of evidence that need to be considered to evaluate
holistic management approaches, upstream that mangroves provide effective mangrove protective attributes (i.e.
management decisions could undermine protection for vulnerable communi- geomorphology and forest intact-
the capacity of mangroves to provide DRR. ties from hazards such as tropical ness), and can therefore potentially
storms and tsunamis and from chronic increase risk by providing a false sense
CONCLUSION AND stressors such as sea-level rise and of security. u
RECOMMENDATIONS coastal erosion.
Although the value of mangroves in pro- • Disaster risk managers can improve
viding DRR services is well-documented outcomes in the use of mangroves for
(albeit with a need for more integrated DRR by carefully considering a range
and multidisciplinary post-disaster stud- of factors in risk reduction analysis,
ies), they are rarely considered in planning planning and management. They References
and development. Even when the role of should consider the viability of cur-
mangroves in DRR is acknowledged, the rent – and potential for future – protec- Algoni, D.M. 2015. The impact of climate
full suite of conditions and issues that tive mangrove ecosystem services in change on mangrove forests. Current
need to be considered to evaluate protec- the context of combined climate and Climate Change Reports, 1(1): 30–39.
tive attributes (e.g. geomorphology and development impacts over time to bet- Blankespoor, B., Dasgupta, S. & Lange, G.M.
forest intactness) may be downplayed ter understand how well and to what 2016. Mangroves as protection from storm
and oversimplified. Along with an over­ extent mangroves might help reduce surges in a changing climate. Washington,
reliance on hard coastal defence systems, risks. Policymakers and managers DC, World Bank.
the lack of fully integrated assessment should understand the interaction of Bunting, P., Rosenqvist, A., Lucas, R., Rebelo,
can inadvertently increase risk by pro- these factors at a landscape scale. L.M., Hilarides, L., Thomas, N., Hardy,
viding a false sense of security among • Integrated, multidisciplinary post- A., Itoh, T., Shimada, M. & Finlayson, C.
coastal communities (Park et al., 2013). disaster studies can increase under- 2018. The global mangrove watch – a new
The consideration of the viability of cur- standing of the DRR efficacy of 2010 global baseline of mangrove extent.
rent – and potential for future – protective mangroves. Remote Sensing, 10(10): 1669.
mangrove ecosystem services, in the face • It is important to consider how long Danda, A.A., Joshi, A., Ghosh, A.K. &
of the combined impacts of change over it will take for mangroves affected by Saha, R. 2017. State of the art report on
time, can aid policymakers and managers extreme events to recover and thereby biodiversity in Indian Sundarbans. New
at a landscape scale. Coastal protection provide their protective services. This Delhi, World Wide Fund for Nature-India.
plans should not rely solely on the pro- should be factored into recovery and Ervin, J., Mulongoy, K.J., Lawrence, K.,
tection benefits of mangroves and related DRR planning. Game, E., Sheppard, D., Bridgewater, P.,

Unasylva 251, Vol. 70, 2019/1


74

Bennett, G., Gidda, S.B. & Bos, P. 2010. tsunami ecological research in India and sites in southwest Florida, USA. Estuaries
Making protected areas relevant: a guide its implications for policy. Environmental and Coasts. DOI: https://doi.org/10.1007/
to integrating protected areas into wider Management, 46(3): 329–339. s12237-019-00564-8
landscapes, seascapes and sectoral plans Munang, R., Thiaw, I., Alverson, K., Liu, Smith, T.J., Anderson, G.H., Balentine, K.,
and strategies. CBD Technical Series No. 44. J. & Han, Z. 2013. The role of ecosystem Tiling, G., Ward, G.A. & Whelan, K.R.
Montreal, Canada, Convention on Biological services in climate change adaptation and 2009. Cumulative impacts of hurricanes on
Diversity. disaster risk reduction. Current Opinion in Florida mangrove ecosystems: sediment
Ezcurra, E., Barrios, E., Ezcurra, P., Environmental Sustainability, 5(1): 47–52. deposition, storm surges and vegetation.
Ezcurra, A., Vanderplank, S., Vidal, Nagelkerken, I.S.J.M., Blaber, S.J.M., Wetlands, 29(1): 24.
O., Villanueva-Almanza, L. & Aburto- Bouillon, S., Green, P., Haywood, M., Spalding, M., McIvor, A., Tonneijck, F.H.,
Oropeza, O. 2019. A natural experiment Kirton, L.G., Meynecke, J.O., Pawlik, Tol, S. & van Eijk, P. 2014. Mangroves
reveals the impact of hydroelectric dams J., Penrose, H.M., Sasekumar, A. & for coastal defence. Guidelines for coastal
on the estuaries of tropical rivers. Science Somerfield, P.J. 2008. The habitat function managers & policy makers. Wetlands
Advances, 5(3): 9875. of mangroves for terrestrial and marine International & The Nature Conservancy.
FAO. 2007. The world’s mangroves 1980– fauna: a review. Aquatic Botany, 89(2): Sutton-Grier, A.E., Wowk, K. & Bamford, H.
2005. FAO Forestry Paper No. 153. Rome. 155–185. 2015. Future of our coasts: the potential for
Grill, G., Lehner, B., Thieme, M., Geenen, Narayan, S., Beck, M.W., Reguero, B.G., natural and hybrid infrastructure to enhance
B., Tickner, D., Antonelli, F., Babu, S., Losada, I. J., Van Wesenbeeck, B., the resilience of our coastal communities,
Borrelli, P., Cheng, L., Crochetiere, H. & Pontee, N., Sanchirico, J.N., Ingram, economies and ecosystems. Environmental
Macedo, H.E. 2019. Mapping the world’s J.C., Lange, G.M. & Burks-Copes, K.A. Science & Policy, 51: 137–148.
free-flowing rivers. Nature, 569(7755): 215. 2016. The effectiveness, costs and coastal Syvitski, J.P., Vörösmarty, C.J., Kettner,
Gupta, H., Kao, S.J. & Dai, M. 2012. The role protection benefits of natural and nature- A.J. & Green, P. 2005. Impact of humans on
of mega dams in reducing sediment fluxes: based defences. PLOS ONE, 11(5): e0154735. the flux of terrestrial sediment to the global
a case study of large Asian rivers. Journal Park, J., Seager, T.P., Rao, P.S.C., Convertino, coastal ocean. Science, 308(5720): 376–380.
of Hydrology, 464: 447–458. M. & Linkov, I. 2013. Integrating risk Vo, Q.T., Kuenzer, C., Vo, Q.M., Moder, F.
Li, X., Liu, J.P., Saito, Y. & Nguyen, V.L. and resilience approaches to catastrophe & Oppelt, N. 2012. Review of valuation
2017. Recent evolution of the Mekong delta management in engineering systems. Risk methods for mangrove ecosystem services.
and the impacts of dams. Earth-Science Analysis, 33(3): 356–367. Ecological Indicators, 23: 431–446.
Reviews, 175: 1–17. Powell, E.J., Tyrrell, M.C., Milliken, A., Ward, R.D., Friess, D.A., Day, R.H. &
Maynard, J., Tracey, D., Williams, G., Tirpak, J.M. & Staudinger, M.D. 2019. Mackenzie, R.A. 2016. Impacts of climate
Andradi-Brown, D.A., Grill, G., Thieme, A review of coastal management approaches change on mangrove ecosystems: a region
M. & Ahmadia, G.N. 2019. Mangrove cover to support the integration of ecological and by region overview. Ecosystem Health and
change between 1996 and 2016 near river- human community planning for climate Sustainability, 2(4): e01211. https://doi.
ocean outlets: a global analysis to identify change. Journal of Coastal Conservation, org/10.1002/ehs2.1211
priority rivers for conservation. Washington, 23(1): 1–18. Zarfl, C., Lumsdon, A.E., Berlekamp, J.,
DC, World Wildlife Fund. DOI: 10.6084/ Radabaugh, K.R., Moyer, R.P., Chappel, Tydecks, L. & Tockner, K. 2015. A global
m9.figshare.8094245 A.R., Dontis, E.E., Russo, C.E., Joyse, boom in hydropower dam construction.
Mukherjee, N., Dahdouh-Guebas, F., K.M., Bownik, M.W., Goeckner, A.H. Aquatic Sciences, 77(1): 161–170. u
Kapoor, V., Arthur, R., Koedam, N., & Khan, N.S. 2019. Mangrove damage,
Sridhar, A. & Shanker, K. 2010. From delayed mortality, and early recovery
bathymetry to bioshields: a review of post- following Hurricane Irma at two landfall

Unasylva 251, Vol. 70, 2019/1


75

FAO FORESTRY

FRA2020 Global Remote Sensing Survey


The FAO Forestry team responsible for the collection of forestry
data from countries and the compilation of the 2020 Global Forest
Resources Assessment (FRA2020) has been conducting training
workshops. The aim of the workshops is to introduce country
experts – who will provide FAO with revised data for FRA2020 – to
new methods for collecting, storing and monitoring information
using satellite imagery and to new tools designed to improve image
processing and interpretation.
The workshops will help develop national expert capacities in the
production of accurate and comparable data following an internation-
© FAO

ally agreed methodology and classification.


Dignitaries gather on stage at the opening of The training sessions and the improved data collection are designed
Asia-Pacific Forestry Week 2019 to assist in the upcoming FRA2020 Global Remote Sensing Survey,
which will create a global dataset for assessing forest area, and
The challenges facing forests in the Asia-Pacific forest-area change, at the regional and global levels.
region Conducted by the FAO Forestry Department with the financial
The 28th session of the Asia-Pacific Forestry Commission (APFC) support of the European Commission, the Global Remote Sensing
was held on 17–21 June 2019 in Incheon, Republic of Korea, at the Survey aims to complement FRA2020 by providing a comprehensive
invitation of the Government of the Republic of Korea. Delegates from overview of global forest resources. It is also intended to empower
22 member countries and 4 United Nations organizations participated national experts through participatory and collaborative approaches
in the session, along with observers and representatives from 21 to develop national capacities in remote sensing assessment.
regional and international intergovernmental and non-governmental The Global Remote Sensing Survey will use the open-source
organizations. software Collect Earth Online, one of FAO’s Open Foris set of tools
In addition to administrative matters, the APFC considered the developed in recent years by the National Aeronautics and Space
following agenda items: the session’s theme of “forests for peace Administration (United States of America) and FAO with Google’s
and well-being”; forest and landscape restoration; community forests, support.
trade and markets; the impacts of technological advances on forests Global Remote Sensing Survey workshops have been held to date
and forestry; FAO’s work on biodiversity; progress in implementing in Argentina, Brazil, China, the Democratic Republic of the Congo,
activities in the region supported by the APFC and FAO; the third India, Madagascar, Paraguay, the Russian Federation and Thailand.
Asia-Pacific Forest Sector Outlook Study (see “Forest Futures” on
p. 78); forests and climate change; the state of forestry in Asia and More information: www.fao.org/forest-resources-assessment
the Pacific; preparations for the 25th session of the Committee
on Forestry and the XV World Forestry Congress; reports and
recommendations from the 2019 Asia-Pacific Forestry Week (see
below); global processes; and implementation of the United Nations
Strategic Plan for Forests and collaboration with the United Nations
Forum on Forests.
The 2019 Asia-Pacific Forestry Week, held in Incheon concurrently
with the 28th session of the APFC, attracted about 2 000 participants
from government, civil society, research, academia and the private
sector. A total of 82 partner events – workshops, seminars and
discussion forums – were convened over the week, generating
rich debate and discussion on a wide range of pressing issues for
forests in the Asia-Pacific region.
Asia-Pacific Forestry Week coincides with sessions of the APFC
to enable dialogue and feedback with the APFC member states
and other forest stakeholders. The event in Incheon was hosted
by FAO and the Korea Forest Service with the support of Incheon
Metropolitan City and 18 institutions acting as stream leaders.

Unasylva 251, Vol. 70, 2019/1


76

WORLD OF FORESTRY

©FAO/LUIS TATO
Beneficiaries of an FAO programme on conservation
agriculture, Lucy Kathegu Kigunda and Gervasio Kigunda,
Two new United Nations "decades" pose on their family farm near Meru, Meru County, Kenya.
FAO and United Nations partner agencies have launched two United Family farming is the focus of a United Nations “decade”
Nations “decades” that will involve considerable contributions by in 2019–2028
FAO – and a significant forestry component – in their implementation:
the United Nations Decade of Family Farming (2019–2028), and the approach is to strengthen the efforts of producer organizations to
United Nations Decade on Ecosystem Restoration (2021–2030). sustain and improve the livelihoods of rural and forest communities,
The United Nations Decade of Family Farming (2019–2028) was ensure fair and equitable access to markets, and encourage the
proclaimed in December 2017 as a way of enabling family farming to inclusion of women as equal partners with men by empowering
transform food systems and play an optimal role in achieving the 2030 them with the same rights and level of participation in community
Agenda for Sustainable Development. FAO and the International Fund decision-making.
for Agricultural Development launched the decade in Rome, Italy, on The United Nations General Assembly officially declared the
29 May 2019. A global action plan was also presented (see “United United Nations Decade on Ecosystem Restoration 2021–2030 on
Nations Decade of Family Farming” on p. 79) to provide guidance on 1 March 2019. Its purpose is to coordinate the restoration of degraded
the steps that need to be taken to achieve the decade’s goal and to ecosystems on a massive scale worldwide to fight the negative
boost support for family farmers, especially in developing countries. consequences of climate change, especially regarding water supply
The forestry component of the United Nations Decade of Family and biodiversity, and to increase food security.
Farming is apparent through the work of the Forest and Farm Facility, It is estimated that ecosystem degradation negatively affects the
a mechanism supported by FAO and partners. The Facility’s main well-being of more than 3 billion people and costs about 10 percent of

Unasylva 251, Vol. 70, 2019/1


77
77

annual gross product in the loss of ecosystem services. Food systems Global Landscapes Forum 2019
and agriculture, the supply of freshwater, protection against hazards Organized alongside the Bonn Climate Change Conference under
and the provision of habitat for pollinators and wildlife are among the United Nations Framework Convention on Climate Change,
the ecosystem services that are declining most rapidly. Ecosystem and building on the momentum on indigenous peoples’ rights built
restoration, on the other hand, can generate trillions of dollars in at the 18th Session of the United Nations Permanent Forum on
ecosystem services and remove large quantities of greenhouse Indigenous Issues in New York on 13–22 April 2019, the Global
gases from the atmosphere. Landscapes Forum (GLF) held in Bonn, Germany, on 23 June 2019
UN Environment and FAO will lead the implementation of the Decade focused on the rights of indigenous peoples and local communities
on Ecosystem Restoration, in collaboration with partners, with the aim and specifically on a rights-based approach to the restoration of
of accelerating existing global restoration goals, such as the Bonn landscapes and forests.
Challenge, the aim of which is to restore 350 million ha of degraded One of the main items on the agenda was the GLF’s contribution to
ecosystems by 2030. the shaping of the United Nations Decade on Ecosystem Restoration.
Among the regional initiatives conducive to that aim are Initiative Participants heard that about 370 million indigenous peoples in
20x20 in Latin America, which aims to restore 20 million ha of 87 countries worldwide manage or have tenure rights to more than
degraded land by 2020, and the AFR100 African Forest Landscape 38 million km2, and they represent a powerful force for protection
Restoration Initiative, which aims to bring 100 million ha of degraded against climate change. When their rights are recognized and
land under restoration by 2030. enforced, indigenous peoples are effective managers of their lands,
Mette Wilkie, the Director of the Forestry Policy and Resources which store vast quantities of carbon and provide habitat for a large
Division at FAO, drew attention to the Action Against Desertification proportion of the world’s biodiversity. The GLF heard that it is impor-
(AAD) programme, a powerful instrument implemented by FAO and tant for the international community to recognize that the relationships
partners funded by the European Union, the aim of which is to promote of indigenous peoples with the natural world are crucial not only for
sustainable land management and restore drylands and degraded the conservation of their own lands but also for the well-being of all.
lands in Africa, the Caribbean and the Pacific. Also at GLF 2019, the Rights and Resources Initiative and the
“Greening the world’s drylands is indeed possible and brings with Indigenous Peoples Major Group for Sustainable Development
it a wealth of associated benefits, from reduction of poverty and presented a first draft of a “gold standard” for rights. The aim of
hunger to mitigation of climate change and reduced risks of conflict,” the standard is to define the principles of secure and proper rights
said Ms Wilkie. that organizations, institutions, governments and the private sector
should apply in the implementation of landscape-based projects,
businesses, initiatives and the law.

Unasylva 251, Vol. 70, 2019/1


78

BOOKS

0.36 cm spine for 72 pg on 90g paper


0.66 cm spine for 112pg on 90g paper

181
181
182

182 FAO FORESTRY PAPER


FAO FORESTRY PAPER

Climate change Climate change


Guide to the classical
biological control of Guide to the classical biological for forest policy-makers for forest policy-makers
An approach for integrating
insect pests in planted control of insect pests in climate change into national forest
An approach for integrating climate change into national
and natural forests policy in support of sustainable forest policy in support of sustainable forest management
planted and natural forests
Guide to the classical biological control of insect pests in planted and natural forests

forest management Version 2.0


Version 2.0
Insect pests damage millions of hectares of forest
worldwide each year. Moreover, the extent of such
damage is increasing as international trade grows,
facilitating the spread of insect pests, and as the The critical role of forests in climate change

Climate change for forest policy-makers


impacts of climate change become more evident. mitigation and adaptation is now widely recognized.
Classical biological control is a well-tried, Forests contribute significantly to climate change
cost-effective approach to the management of mitigation through their carbon sink and carbon
invasive forest pests. It involves the importing of storage functions. They play an essential role in
“natural enemies” of non-native pests from their reducing vulnerabilities and enhancing adaptation
countries of origin with the aim of establishing of people and ecosystems to climate change and
permanent, self-sustaining populations capable of climate variability, the negative impacts of which
sustainably reducing pest populations below are becoming increasingly evident in many parts of
damaging levels. the world.

A great deal of knowledge on classical biological In many countries climate change issues have not
control has been accumulated worldwide in the last been fully addressed in national forest policies,
few decades. This publication, which was written by forestry mitigation and adaptation needs at national
a team of experts, distils that information in a clear, level have not been thoroughly considered in
concise guide aimed at helping forest-health national climate change strategies, and cross-sectoral
practitioners and forest managers – especially in dimensions of climate change impacts and response
developing countries – to implement successful measures have not been fully appreciated. This
classical biological control programmes. It provides publication seeks to provide a practical approach to
general theory and practical guidelines, explains the the process of integrating climate change into
“why” and “how” of classical biological control in national forest programmes. The aim is to assist
forestry, and addresses the potential risks associated senior officials in government administrations and
with such programmes. It features 11 case studies of the representatives of other stakeholders, including
successful efforts worldwide to implement classical civil society organizations and the private sector,
biological control. prepare the forest sector for the challenges and
opportunities posed by climate change.

This document complements a set of guidelines


prepared by FAO in 2013 to support forest managers
incorporate climate change considerations into
forest management plans and practices.

ISBN 978-92-5-131335-0 ISSN 0258-6150 ISBN 978-92-5-131094-6 ISSN 0258-6150


ISSN 0258-6150

ISSN 0258-6150
FAO FAO
FORESTRY FORESTRY
PAPER PAPER
FAO

FAO
9 7 8 9 2 5 1 3 1 3 3 5 0
9 7 8 9 2 5 1 3 1 0 9 4 6
CA3677EN/1/03.19
CA2309EN/1/11.18

182 181

Managing invasive forest pests with classical The role of forests in climate-change mitigation
biological control Climate change for forest policy-makers: an approach for integrating climate change

Guide to the classical biological control of insect pests in planted and natural into national forest policy in support of sustainable forest management. Version 2.0.

forests. FAO Forestry Paper No. 182. M. Kenis, B.P. Hurley, F. Colombari, FAO Forestry Paper No. 181. 2018. Rome, FAO. ISBN 978-92-5-131094-6.

S. Lawson, J. Sun, C. Wilcken, R. Weeks, & S. Sathyapala. 2019. Rome, FAO. Forests contribute significantly to climate-change mitigation through
ISBN 978‑92‑5‑131335‑0. their carbon sink and carbon storage functions. They play essential
Insect pests damage millions of hectares of forest each year roles in reducing vulnerabilities and enhancing the adaptation of
worldwide. Moreover, the extent of such damage is increasing as people and ecosystems to climate change and climate variability,
international trade grows (facilitating the spread of insect pests) and the negative impacts of which are becoming increasingly evident in
as the impacts of climate change become more evident. many parts of the world.
Classical biological control is a well-tried, cost-effective approach In many countries, however, climate change is not being fully
to the management of invasive forest pests. It involves the importing addressed in national forest policies. Moreover, forest-related
of “natural enemies” of non-native pests from their countries of origin climate-change mitigation and adaptation needs have not been
with the aim of establishing permanent, self-sustaining populations thoroughly considered in national climate-change strategies, and
capable of sustainably reducing pest populations below damaging the cross-sectoral dimensions of climate-change impacts and
levels. A great deal of knowledge on classical biological control has responses are underappreciated. This publication provides a
been accumulated worldwide in the last few decades. practical approach to the process of integrating climate change into
This publication, which was written by a team of experts, distils national forest programmes. The aim is to assist senior officials in
such knowledge in a clear, concise guide aimed at helping forest- government administrations and the representatives of other stake-
health practitioners and forest managers – especially in developing holders, including civil-society organizations and the private sector,
countries – to implement successful classical biological control to prepare the forest sector for the challenges and opportunities
programmes. It provides general theory and practical guidelines, posed by climate change.
explains the “why” and “how” of classical biological control in This document complements a set of guidelines prepared by FAO
forestry, and addresses the potential risks associated with such in 2013 to support forest managers in incorporating climate-change
programmes. It features 11 case studies of successful efforts to considerations into forest management plans and practices.
implement classical biological control.
Available online: www.fao.org/3/CA2309EN/ca2309en.pdf
Available online: www.fao.org/3/ca3677en/CA3677EN.pdf

Unasylva 251, Vol. 70, 2019/1


79

0.66 cm spine for 132 pg on 90g paper


180

180
FAO FORESTRY PAPER

Making forest concessions


in the tropics work to Making forest concessions in FOREST FUTURES
achieve the 2030 Agenda: the tropics work to achieve the Sustainable pathways for forests, landscapes
Making forest concessions in the tropics work to achieve the 2030 Agenda: Voluntary Guidelines

Voluntary Guidelines and people in the Asia-Pacific region


2030 Agenda: Voluntary Guidelines
Sustainable wood products and their value chains can
play a fundamental role in achieving the objectives
stated in the 2030 Agenda and the Paris Agreement,
delivering a wide range of benefits to populations in
remote forest areas as well as to local, regional and
global society. Generation of income and employment,
disaster risk reduction, and reduction of the material
and carbon footprint of the planet are some of the
direct contributions sustainable forest products can
provide to the SDGs and the climate change
commitments. Furthermore, sustainable management
of natural forests reduces forest degradation and
forest production can increase the opportunity cost
for deforestation, while generating revenues for
conservation strategies.

These Voluntary Guidelines for forest concessions


focus on concessions as a forest policy instrument for
the delivery of sustainable forest management in the
tropics, building on lessons learned from success and
failures in implementing forest concession. The
guidelines offer a practical participatory management
approach to support forest concession regimes to be
reliable sources of sustainable wood and non-wood
forest products and contribute to realizing the full
contribution of forestry to the 2030 Agenda.

With the financial support of

ISBN 978-92-5-130547-8 ISSN 0251-6150


ISSN 0258-6150

FAO
FORESTRY
PAPER
Asia-Pacific Forest Sector Outlook Study III
FAO

9 7 8 9 2 5 1 3 0 5 4 7 8
I9487EN/1/05.18

180

Concessions for sustainable wood production in The future of forests in Asia and the Pacific
the tropics Forest futures: sustainable pathways for forests, landscapes and people in the Asia-

Making forest concessions in the tropics work to achieve the 2030 Agenda: Pacific region. Asia-Pacific Forest Sector Outlook Study III. 2019. Bangkok, FAO.

Voluntary guidelines. FAO Forestry Paper No. 180. Y.T. Tegegne, J. Van Brusselen, ISBN 978-92-5-131457-9.

M. Cramm, T. Linhares-Juvenal, P. Pacheco, C. Sabogal & D. Tuomasjukka. 2018. Forests and landscapes in the Asia-Pacific region are under
Rome, FAO and the European Forest Institute. ISBN 978-92-5-130547-8. increasing pressure from economic development, climate change,
Sustainable wood products and their value chains can play demographic shifts, conflicts over tenure and land use, and other
fundamental roles in achieving the objectives of the 2030 Agenda stressors. This publication, the third Asia-Pacific Forest Sector
for Sustainable Development and the Paris Agreement on climate Outlook Study, presents scenarios and a strategic analysis to help
change, delivering a wide range of benefits to communities in policymakers and other actors understand the implications of these
remote forest areas as well as to local, regional and global societies. stressors for forests and forestry in the Asia-Pacific region and how
Sustainable forest products can make direct contributions to the best to address the challenges ahead.
Sustainable Development Goals, such as by providing income The product of outstanding collaboration among institutions,
and employment, abating the risk of disasters, and reducing networks and more than 800 individuals across the region, the
environmental footprints. Moreover, the sustainable management of study examines the drivers of change in the region’s forest sector
natural forests reduces forest degradation, while sustainable forest and explores three scenarios – business-as-usual, aspirational and
production can increase the opportunity costs of deforestation and disruptive – to 2030 and 2050. It shows that “more of the same” will
generate revenues for conservation strategies. likely lead to highly negative outcomes over both time horizons.
These voluntary guidelines for forest concessions focus on On the other hand, the adoption of landscape approaches and
concessions as a policy instrument for the delivery of sustainable other key measures could help realize the enormous potential of
forest management in the tropics, building on lessons learned from forests – with their capacity to simultaneously perform multiple
successes and failures in implementing forest concessions. The economic, social and environmental functions – to help achieve
guidelines offer a practical participatory management approach to development goals in and beyond the forest sector. A key message
ensuring that forest concession regimes act as reliable sources of of the report is that the region must respond now to ensure the
sustainable wood and non-wood forest products and contribute to resilience of forests, landscapes and communities and thereby avoid
realizing the full contributions of forestry to the 2030 Agenda for catastrophic outcomes. The report sets out seven “robust actions”
Sustainable Development. for operationalizing this response.

Available online: www.fao.org/3/I9487EN/i9487en.pdf Available online: www.fao.org/3/ca4627en/ca4627en.pdf

Unasylva 251, Vol. 70, 2019/1


80

UNITED NATIONS DECADE OF


FAMILY FARMING 2019-2028
Global Action Plan

Joint Secretariat of the UN Decade of


Family Farming

Food and Agriculture Organization of the


United Nations (FAO)
Viale delle Terme di Caracalla
00153 Rome, Italy

Decade-Of-Family-Farming-Secretariat@fao.org
www.fao.org/family-farming-decade
#FamilyFarming
UNITED NATIONS DECADE OF
FAMILY FARMING 2019-2028
ISBN 978-92-5-131472-2

Global Action Plan


9 7 8 9 2 5 1 3 1 4 7 2 2 1
CA4672EN/1/05.19

Strengthening the contributions of family farming to Improving the governance of forest tenure
food systems Assessing the governance of tenure for improving forests and livelihoods: a tool

United Nations Decade of Family Farming 2019-2028: Global Action Plan. to support the implementation of the Voluntary Guidelines on the Responsible

2019. Rome, FAO and International Fund for Agricultural Development (IFAD). Governance of Tenure. FAO Forestry Working Paper No. 13. 2019. Rome, FAO.

ISBN 978‑92-5-131472-2. ISBN 978-92-5-131553-8.

By placing family farming at the centre of the international agenda Governments around the world have been attempting for many
for a period of ten years, the United Nations Decade of Family years to strengthen and give formal recognition to customary tenure.
Farming (2019–2028) provides an unprecedented opportunity to In addition, forestry departments have introduced various types of
achieve positive change in food systems globally. Family farmers participatory arrangements recognizing certain resource-use rights of
have proven their capacity to develop new strategies and provide local communities with the purpose of improving forest governance
innovative responses to emerging economic, social and environmental and reducing poverty.
challenges. They don’t just produce food: they simultaneously perform This assessment tool was developed to better understand the
environmental, social and cultural functions, act as custodians of strengths and limitations of such forest-tenure reforms. It uses the
biodiversity, and help preserve landscapes and maintain community internationally endorsed Voluntary Guidelines on the Responsible
and cultural heritage. Family farmers also have the knowledge to Governance of Tenure of Land, Fisheries and Forests as its basis.
produce nutritious and culturally appropriate food as part of local Although the tool enables the assessment of all forms of tenure
traditions. arrangements, it can be particularly helpful for assessing those that
The Global Action Plan of the United Nations Decade of Family recognize customary tenure in forestry through participatory for-
Farming (2019–2028) represents the tangible result of an exten- estry initiatives such as collaborative forestry, community forestry
sive and inclusive global consultation process involving diverse and smallholder forestry. The tool also enables the identification
partners worldwide. The aim is to mobilize coordinated actions to and assessment of customary-tenure systems not recognized in
help family farmers overcome the challenges they face, increase statutory law.
their investment capacity, and thereby obtain the benefits of their As experienced in several test countries, the findings and recom-
contributions in transforming societies and putting in place long- mendations emerging from the assessment of tenure arrangements
term, sustainable solutions. can provide valuable insights into the strengths and limitations of
existing arrangements and reforms and help generate ideas for
Available online: www.fao.org/3/ca4672en/ca4672en.pdf improving their performance in forest governance, strengthening
local livelihoods and contributing to the Sustainable Development
Goals.

Available online: www.fao.org/3/ca5039en/CA5039EN.pdf

Unasylva 251, Vol. 70, 2019/1


81
81

Valuing forest
ecosystem services
A training manual
for planners and project developers

FORESTRY
WORKING
PAPER

11

Assessing the benefits of community forestry How to correctly estimate forest ecosystem services
A framework to assess the extent and effectiveness of community-based forestry. Valuing forest ecosystem services: a training manual for planners and project

FAO Forestry Working Paper No. 12. 2019. Rome, FAO. ISBN 978-92-5-131547-7. developers. FAO Forestry Working Paper No. 11. M. Masiero, D. Pettenella,

There has been significant expansion in the area under community- M. Boscolo, S.K. Barua, I. Animon & J.R. Matta. 2019. Rome, FAO. ISBN 978-92-

based forestry (CBF) in the last four decades, involving a broad array 5-131215-5.

of initiatives that favour people’s participation in forestry. The failure to appropriately consider the full economic value
This aim of this assessment framework is to help in generating of ecosystem services in decision-making contributes to the
insights into the successes and shortcomings of CBF at the country continued degradation and loss of ecosystems and biodiversity.
level. The framework can also provide a means for determining Most ecosystem services are considered public goods and tend
and tracking the extent and effectiveness of the broad spectrum to be overexploited by society. Recognizing, demonstrating and
of CBF initiatives. capturing the value of ecosystem services, on the other hand, can
Well-performing CBF has the potential to rapidly restore forests help in setting policy directions for ecosystem management and
in ecological terms and scale up sustainable forest management conservation and thus in increasing the provision of ecosystem
to the national level while improving livelihoods for many millions of services and their contributions to human well-being.
marginalized people worldwide. In so doing, CBF has the potential The aim of this manual is to increase understanding of eco­
to contribute significantly to the Sustainable Development Goals. system services and their valuation. The target audience comprises
government officers in planning units and field-level officers and prac-
Available online: www.fao.org/3/ca4987en/CA4987EN.pdf titioners in key government departments in Bangladesh responsible
for project development, including the Ministry of Environment and
Forests and its agencies. Most of the examples and case studies
presented herein, therefore, are tailored to the Bangladeshi con-
text, but the general concepts, approaches and methods can be
applied to a broad spectrum of situations. This manual focuses on
valuing forest-related ecosystem services, including those provided
by trees outside forests. It is expected to improve valuation efforts
and help ensure the better use of such values in policymaking and
decision-making.
Among other things, the manual explores the basics of financial
mathematics (e.g. the time value of money; discounting; cost–benefit
analysis; and profitability and risk indicators); the main methods of
economic valuation; examples of the valuation of selected eco­system
services; and inputs for considering values in decision-making.

Available online: www.fao.org/3/ca2886en/CA2886EN.pdf

Unasylva 251, Vol. 70, 2019/1


82
82

BOOKS
8
Agroforestry and tenure

Agroforestry and tenure


State of
Mediterranean
Forests 2018

For more information, please contact:

Forestry Department
E-mail: fo-library@fao.org ISBN 978-92-5-131467-8

Web address: www.fao.org/forestry/en


FORESTRY
Food and Agriculture Organization of the United Nations WORKING
9 7 8 9 2 5 1 3 1 4 6 7 8
Viale delle Terme di Caracalla PAPER
FAO

CA4662EN/1/05.19

8
00153 Rome, Italy

Addressing tenure-related challenges in Climate change – opportunities and risks for


agroforestry Mediterranean forests
Agroforestry and tenure. FAO Forestry Working Paper No. 8. S. Borelli, E. Simelton, State of Mediterranean forests 2018. 2018. Rome and Marseille, France, FAO and

S. Aggarwal, A. Olivier, M. Conigliaro, A. Hillbrand, D. Garant & H. Desmyttere. Plan Bleu. ISBN 978-92-5-131047-2 (FAO).

2019. Rome, FAO and World Agroforestry (ICRAF). ISBN 978-92-5-131467-8. The Mediterranean region has more than 25 million ha of forests
Agroforestry is gaining new ground in the quest for climate-smart and about 50 million ha of other wooded lands. They make vital
agricultural practices due to its potential to sequester carbon and contributions to rural development, poverty alleviation and food
mitigate climate change while increasing the socio-economic and security and to the agriculture, water, tourism and energy sectors.
environmental sustainability of rural development. Changes in climate, societies and lifestyles in the Mediterranean,
Yet agroforestry continues to face challenges, such as unfavour- however, could have serious negative consequences for forests.
able policy incentives, legal constraints, and poor coordination Both to compensate for the lack of data on Mediterranean for-
among sectors. In particular, many agroforestry researchers and ests and to provide a sound basis for their future management, the
practitioners have identified insecure land and resource tenure as Committee on Mediterranean Forestry Questions-Silva Mediterranea
a major obstacle to the promotion of this practice. requested FAO, in collaboration with other institutions, to prepare
This publication reviews the main tenure-related challenges that and regularly update a report on the state of Mediterranean for-
can affect agroforestry adoption with the aim of informing policies ests. The first edition of State of Mediterranean Forests, published
and project implementation. Challenges include tenure insecurity in 2013, has become an important reference work. The aim of this
– regarding either land or its products – that undermines the adop- second edition is to demonstrate the importance of Mediterranean
tion of agroforestry; small plot sizes; policies limiting access to and forests in tackling issues of global significance, such as climate
the use of land by women and minority groups; and the barriers change and population growth. Mediterranean forests also have
presented by some customary regimes. a role to play in helping countries meet their international commit-
Drawing on practical case studies, the publication presents ments on forests, particularly the Sustainable Development Goals
measures and approaches that could help drive the adoption of and the objectives of the three Rio Conventions.
agroforestry. It concludes with recommendations for formulating
and implementing tenure policies that promote agroforestry. Available online: www.fao.org/3/ca2081en/CA2081EN.pdf
State of Mediterranean Forests 2018 – executive summary is also
Available online: www.fao.org/3/CA4662en/CA4662en.pdf available in
English: www.fao.org/3/ca3759en/CA3759EN.pdf
French: www.fao.org/3/ca3759fr/CA3759FR.pdf

Unasylva 251, Vol. 70, 2019/1


83
83

7
Small-scale forest enterprises
in Latin America The role of forest producer
Unlocking their potential for sustainable livelihoods organizations in social protection

The role of forest producer organizations in social protection


For more information, please contact:

Qiang Ma
Forestry Officer
Forestry Department, FAO
Viale delle Terme di Caracalla
00153 Rome, Italy ISBN 978-92-5-130789-2

Email: Qiang.Ma@fao.org
FORESTRY
WORKING Website: www.fao.org/forestry FORESTRY
PAPER WORKING
9 7 8 9 2 5 1 3 0 7 8 9 2
PAPER

10

FAO
CA0370EN/1/07.18

Latin America’s vital small-scale forest enterprises Extending social protection through forest producer
Small-scale forest enterprises in Latin America: unlocking their potential for organizations
sustainable livelihoods. FAO Forestry Working Paper No. 10. F. Del Gatto, J. The role of forest producer organizations in social protection. FAO Forestry Working

Mbairamadji, M. Richards & D. Reeb. 2018. Rome, FAO. ISBN 978-92-5-131119-6. Paper No. 7. N. Tirivayi, L. Nennen & W. Tesfaye. 2018. Rome, FAO. ISBN 978-92-5-

Latin America has a rich and unique experience in the development 130789-2.

of small-scale forest enterprises (SSFEs). Mexico, which has had a This study identifies a broad range of factors that can enable or
vigorous SSFE sector since the 1970s, was a pioneer, and several constrain the provision of social protection benefits by forest producer
other countries followed suit in subsequent decades. SSFEs are organizations (FPOs). Enabling factors include secure land rights,
now numerous in many Latin American countries, and some have robust leadership and management, revenues and market accessibility
developed strong associations and alliances to promote and sustain for forest resources, a supportive institutional environment, and a
their growth. Nevertheless, the potential of SSFEs is yet to be fully favourable social and political context in communities.
realized. FPOs have opportunities to obtain financial and technical assis-
This publication focuses on SSFE development in Latin America. It tance that can boost their viability and thereby create fiscal space
documents their status and recent trends, identifies key challenges for the provision of social protection. The collective and participa-
and opportunities, and presents recommendations for strengthening tory nature of FPOs is an asset for implementing social protection
them and their role in sustainable development. benefits. International climate-change initiatives provide potential
avenues for strengthening and supporting FPOs in the provision of
Available online: www.fao.org/3/ca2431en/CA2431EN.pdf these benefits. Nevertheless, FPOs need to overcome an array of
constraints that arise from their geographical remoteness, climate
variability, insecure tenure, poor access to credit and finance, con-
flict, and social and political exclusion.

Available online: www.fao.org/3/CA0370EN/ca0370en.pdf

Unasylva 251, Vol. 70, 2019/1


84
84

BOOKS

Analyse widely,
act deeply
Forest and farm producer
organisations and the goal of
climate resilient landscapes
James Mayers

Discussion Paper Forests

Keywords:
April 2019 Forests, Finance, Forest and
Farm Facility

The contributions of producer organizations to


climate resilience
Analyse widely, act deeply: forest and farm producer organisations and the goal

of climate resilient landscapes. IIED Discussion Paper. J. Mayers. 2019. London,

International Institute for Environment and Development (IIED). ISBN 978-1-78431-

681-5.

Local organizations, thriving among smallholders who are dependent


on adjacent forests or trees growing on their farms, constitute perhaps
the world’s biggest and most effective force for improving rural
livelihoods and sustainability. They face fast-changing pressures,
however.
Many forest and farm producer organizations are likely to find
it useful to have an organizational goal of contributing to climate-
resilient landscapes. Various international programmes can help
in understanding and supporting such contributions – especially
through practical actions for climate adaptation and mitigation, and
forest restoration. “Landscape approaches” are helpful for analys-
ing the various connected issues, while context-specific, politically
savvy planning is needed for effective action.
This paper explores the possible motivations and actions for
climate-resilient landscapes in four sorts of forest and farm
producer organizations: indigenous peoples’ organizations;
community forest organizations; forest and farm producer groups;
and processing groups in urban and peri-urban contexts.

Available online: https://pubs.iied.org/pdfs/13610IIED.pdf

Unasylva 251, Vol. 70, 2019/1


Electronic subscription to
Unasylva
Would you like to continue receiving a hard copy of
250
ISSN 0041-6436
Unasylva or would you rather receive it electronically –
or perhaps both?
If you wish to substitute your hard-copy subscription for
An international journal of forestry and forest industries Vol. 69 2018/1

FORESTS AND electronic-only, please write to Unasylva@fao.org with


SUSTAINABLE
CITIES “Electronic subscription only” in the subject line.
If you wish to receive both a hard copy and an electronic
copy, please write to Unasylva@fao.org with “Electronic
and hard-copy subscription” in the subject line.
Please provide the relevant contact details in the email.
Unasylva will continue to be a free-subscription journal
available in English, French and Spanish.

A guide to participatory learning


on forests and water

ADVANCING THE
FOREST AND WATER NEXUS
A CAPACITY DEVELOPMENT FACILITATION GUIDE The FFAO Forest and Water Programme has
produced this module-based facilitation guide to
prod
provide background information, resource
prov
materials and a facilitation plan to support the
mate
delivery of participatory learning workshops and
deliv
capacity-development
capa programmes on the
forest–water nexus.
fores
The guide
g can be downloaded free of charge at
www.fao.org/documents/card/en/c/ca6483en
www
ISBN 978-92-5-131910-9 ISSN 0041-6436

Unasylva@fao.org 9 789251 319109

www.fao.org/forestry/unasylva
CA6842EN/1/11.19

You might also like