Special Report: Special Report on Climate Change and Land
Ch 03

Desertification

Coordinating Lead Authors

  • Alisher Mirzabaev (Germany, Uzbekistan)
  • Jianguo Wu (China)

Lead Authors

  • Jason Evans (Australia)
  • Felipe Garcia Oliva (Mexico)
  • Ismail Abdel Galil Hussein (Egypt)
  • Muhammad Mohsin Iqbal (Pakistan)
  • Joyce Kimutai (Kenya)
  • Tony Knowles (South Africa)
  • Francisco Meza (Chile)
  • Dalila Nedjraoui (Algeria)
  • Fasil Tena (Ethiopia)
  • Murat Türkeş (Turkey)
  • Ranses José Vázquez (Cuba)
  • Mark Weltz (United States)

Contributing Authors

  • Mansour Almazroui (Saudi Arabia)
  • Hamda Aloui (Tunisia)
  • Hesham El-Askary (Egypt)
  • Abdul Rasul Awan (Pakistan)
  • Céline Bellard (France)
  • Arden Burrell (Australia)
  • Stefan van der Esch (Netherlands)
  • Robyn Hetem (South Africa)
  • Kathleen Hermans (Germany)
  • Margot Hurlbert (Canada)
  • Jagdish Krishnaswamy (India)
  • Zaneta Kubik (Poland)
  • German Kust (Russia)
  • Eike Lüdeling (Germany)
  • Johan Meijer (Netherlands)
  • Ali Mohammed (Egypt)
  • Katerina Michaelides (Cyprus, United Kingdom)
  • Lindsay Stringer (United Kingdom)
  • Stefan Martin Strohmeier (Austria)
  • Grace Villamor (Philippines)

Review Editors

  • Mariam Akhtar-Schuster (Germany)
  • Fatima Driouech (Morocco)
  • Mahesh Sankaran (India)

Chapter Scientists

  • Chuck Chuan Ng (Malaysia)
  • Helen Berga Paulos (Ethiopia)

FAQ 3.1 | How does climate change affect desertification?

Desertification is land degradation in drylands. Climate change and desertification have strong interactions. Desertification affects climate change through loss of fertile soil and vegetation. Soils contain large amounts of carbon, some of which could be released to the atmosphere due to desertification, with important repercussions for the global climate system. The impacts of climate change on desertification are complex and knowledge on the subject is still insufficient. On the one hand, some dryland regions will receive less rainfall and increases in temperatures can reduce soil moisture, harming plant growth. On the other hand, the increase of CO2 in the atmosphere can enhance plant growth if there are enough water and soil nutrients available.

FAQ 3.2 | How can climate change induced desertification be avoided, reduced or reversed?

Managing land sustainably can help avoid, reduce or reverse desertification, and contribute to climate change mitigation and adaptation. Such sustainable land management practices include reducing soil tillage and maintaining plant residues to keep soils covered, planting trees on degraded lands, growing a wider variety of crops, applying efficient irrigation methods, improving rangeland grazing by livestock and many others.

FAQ 3.3 | How do sustainable land management practices affect ecosystem services and biodiversity?

Sustainable land management practices help improve ecosystems services and protect biodiversity. For example, conservation agriculture and better rangeland management can increase the production of food and fibres. Planting trees on degraded lands can improve soil fertility and fix carbon in soils. Sustainable land management practices also support biodiversity through habitat protection. Biodiversity protection allows for the safeguarding of precious genetic resources, thus contributing to human well-being.

Figure 3.1
View details
Figure 3.2
View details
Figure 3.3
View details
Figure 3.4
View details
Figure 3.5
View details
Figure 3.6
View details
Figure 3.7
View details
Figure 3.8
View details
Figure 3.9
View details
Figure 3.10
View details
Figure 3.11a
View details
Figure 3.11b
View details
Figure 3.11c
View details
Figure 3.11d
View details
Figure 3.12b
View details
Figure 3.13
View details
Figure 3.14
View details
Figure 3.15a
View details
Figure 3.15b
View details
Figure 3.15c
View details
Figure-3.15d
View details
Figure 3.16
View details
Figure 3.17a
View details
Figure-3.17b
View details
Figure 3.17c
View details
Figure 3.18
View details
Figure 3.19a
View details
Figure 3.19b
View details
ES

Executive Summary

Desertification is land degradation in arid, semi-arid, and dry sub-humid areas, collectively known as drylands, resulting from many factors, including human activities and climatic variations. The range and intensity of desertification have increased in some dryland areas over the past several decades (high confidence). Drylands currently cover about 46.2% (±0.8%) of the global land area and are home to 3 billion people. The multiplicity and complexity of the processes of desertification make its quantification difficult. Desertification hotspots, as identified by a decline in vegetation productivity between the 1980s and 2000s, extended to about 9.2% of drylands (±0.5%), affecting about 500 (±120) million people in 2015. The highest numbers of people affected are in South and East Asia, the circum Sahara region including North Africa and the Middle East including the Arabian Peninsula (low confidence). Other dryland regions have also experienced desertification. Desertification has already reduced agricultural productivity and incomes (high confidence) and contributed to the loss of biodiversity in some dryland regions (medium confidence). In many dryland areas, spread of invasive plants has led to losses in ecosystem services (high confidence), while over-extraction is leading to groundwater depletion (high confidence). Unsustainable land management, particularly when coupled with droughts, has contributed to higher dust-storm activity, reducing human well-being in drylands and beyond (high confidence). Dust storms were associated with global cardiopulmonary mortality of about 402,000 people in 2005. Higher intensity of sand storms and sand dune movements are causing disruption and damage to transportation and solar and wind energy harvesting infrastructures (high confidence). {3.1.1, 3.1.4, 3.2.1, 3.3.1, 3.4.1, 3.4.2, 3.4.2, 3.7.3, 3.7.4}

Attribution of desertification to climate variability and change, and to human activities, varies in space and time (high confidence). Climate variability and anthropogenic climate change, particularly through increases in both land surface air temperature and evapotranspiration, and decreases in precipitation, are likely to have played a role, in interaction with human activities, in causing desertification in some dryland areas. The major human drivers of desertification interacting with climate change are expansion of croplands, unsustainable land management practices and increased pressure on land from population and income growth. Poverty is limiting both capacities to adapt to climate change and availability of financial resources to invest in sustainable land management (SLM) (high confidence). {3.1.4, 3.2.2, 3.4.2}

Climate change will exacerbate several desertification processes (medium confidence). Although CO2 fertilisation effect is enhancing vegetation productivity in drylands (high confidence), decreases in water availability have a larger effect than CO2 fertilisation in many dryland areas. There is high confidence that aridity will increase in some places, but no evidence for a projected global trend in dryland aridity (medium confidence). The area at risk of salinisation is projected to increase in the future (limited evidence, high agreement). Future climate change is projected to increase the potential for water driven soil erosion in many dryland areas (medium confidence), leading to soil organic carbon decline in some dryland areas. {3.1.1, 3.2.2, 3.5.1, 3.5.2, 3.7.1, 3.7.3}

Risks from desertification are projected to increase due to climate change (high confidence). Under shared socio-economic pathway SSP2 (‘Middle of the Road’) at 1.5°C, 2°C and 3°C of global warming, the number of dryland population exposed (vulnerable) to various impacts related to water, energy and land sectors (e.g., water stress, drought intensity, habitat degradation) is projected to reach 951 (178) million, 1152 (220) million and 1285 (277) million, respectively. While at global warming of 2°C, under SSP1 (‘Sustainability’), the exposed (vulnerable) dryland population is 974 (35) million, and under SSP3 (‘Fragmented World’) it is 1267 (522) million. Around half of the vulnerable population is in South Asia, followed by Central Asia, West Africa and East Asia. {2.2, 3.1.1, 3.2.2, 3.5.1, 3.5.2, 7.2.2}

Desertification and climate change, both individually and in combination, will reduce the provision of dryland ecosystem services and lower ecosystem health, including losses in biodiversity (high confidence). Desertification and changing climate are projected to cause reductions in crop and livestock productivity (high confidence), modify the composition of plant species and reduce biological diversity across drylands (medium confidence). Rising CO2 levels will favour more rapid expansion of some invasive plant species in some regions. A reduction in the quality and quantity of resources available to herbivores can have knock-on consequences for predators, which can potentially lead to disruptive ecological cascades (limited evidence, low agreement). Projected increases in temperature and the severity of drought events across some dryland areas can increase chances of wildfire occurrence (medium confidence). {3.1.4, 3.4.1, 3.5.2, 3.7.3}

Increasing human pressures on land, combined with climate change, will reduce the resilience of dryland populations and constrain their adaptive capacities (medium confidence). The combination of pressures coming from climate variability, anthropogenic climate change and desertification will contribute to poverty, food insecurity, and increased disease burden (high confidence), as well as potentially to conflicts (low confidence). Although strong impacts of climate change on migration in dryland areas are disputed (medium evidence, low agreement), in some places, desertification under changing climate can provide an added incentive to migrate (medium confidence). Women will be impacted more than men by environmental degradation, particularly in those areas with higher dependence on agricultural livelihoods (medium evidence, high agreement). {3.4.2, 3.6.2}

Desertification exacerbates climate change through several mechanisms such as changes in vegetation cover, sand and dust aerosols and greenhouse gas fluxes (high confidence). The extent of areas in which dryness (rather than temperature) controls CO2 exchange has increased by 6% between 1948 and 2012, and is projected to increase by at least another 8% by 2050 if the expansion continues at the same rate. In these areas, net carbon uptake is about 27% lower than in other areas (low confidence). Desertification also tends to increase albedo, decreasing the energy available at the surface and associated surface temperatures, producing a negative feedback on climate change (high confidence). Through its effect on vegetation and soils, desertification changes the absorption and release of associated greenhouse gases (GHGs). Vegetation loss and drying of surface cover due to desertification increases the frequency of dust storms (high confidence). Arid ecosystems could be an important global carbon sink, depending on soil water availability (medium evidence, high agreement). {3.3.3, 3.4.1, 3.5.2}

Site and regionally-specific technological solutions, based both on new scientific innovations and indigenous and local knowledge (ILK), are available to avoid, reduce and reverse desertification, simultaneously contributing to climate change mitigation and adaptation (high confidence). SLM practices in drylands increase agricultural productivity and contribute to climate change adaptation with mitigation co-benefits (high confidence). Integrated crop, soil and water management measures can be employed to reduce soil degradation and increase the resilience of agricultural production systems to the impacts of climate change (high confidence). These measures include crop diversification and adoption of drought-resilient econogically appropriate plants, reduced tillage, adoption of improved irrigation techniques (e.g., drip irrigation) and moisture conservation methods (e.g., rainwater harvesting using indigenous and local practices), and maintaining vegetation and mulch cover. Conservation agriculture increases the capacity of agricultural households to adapt to climate change (high confidence) and can lead to increases in soil organic carbon over time, with quantitative estimates of the rates of carbon sequestration in drylands following changes in agricultural practices ranging between 0.04 and 0.4 t ha–1 (medium confidence). Rangeland management systems based on sustainable grazing and re-vegetation increase rangeland productivity and the flow of ecosystem services (high confidence). The combined use of salt-tolerant crops, improved irrigation practices, chemical remediation measures and appropriate mulch and compost is effective in reducing the impact of secondary salinisation (medium confidence). Application of sand dune stabilisation techniques contributes to reducing sand and dust storms (high confidence). Agroforestry practices and shelterbelts help reduce soil erosion and sequester carbon. Afforestation programmes aimed at creating windbreaks in the form of ‘green walls’ and ‘green dams’ can help stabilise and reduce dust storms, avert wind erosion, and serve as carbon sinks, particularly when done with locally adapted native and other climate resilient tree species (high confidence). {3.4.2, 3.6.1, 3.7.2}

Investments into SLM, land restoration and rehabilitation in dryland areas have positive economic returns (high confidence). Each USD invested into land restoration can have social returns of about 3–6 USD over a 30-year period. Most SLM practices can become financially profitable within 3 to 10 years (medium evidence, high agreement). Despite their benefits in addressing desertification, mitigating and adapting to climate change, and increasing food and economic security, many SLM practices are not widely adopted due to insecure land tenure, lack of access to credit and agricultural advisory services, and insufficient incentives for private land-users (robust evidence, high agreement). {3.6.3}

ILK often contributes to enhancing resilience against climate change and combating desertification (medium confidence). Dryland populations have developed traditional agroecological practices which are well adapted to resource-sparse dryland environments. However, there is robust evidence documenting losses of traditional agroecological knowledge. Traditional agroecological practices are also increasingly unable to cope with growing demand for food. Combined use of ILK and new SLM technologies can contribute to raising the resilience to the challenges of climate change and desertification (high confidence). {3.1.3, 3.6.1, 3.6.2}

Policy frameworks promoting the adoption of SLM solutions contribute to addressing desertification as well as mitigating and adapting to climate change, with co-benefits for poverty eradication and food security among dryland populations (high confidence). Implementation of Land Degradation Neutrality policies allows populations to avoid, reduce and reverse desertification, thus contributing to climate change adaptation with mitigation co-benefits (high confidence). Strengthening land tenure security is a major factor contributing to the adoption of soil conservation measures in croplands (high confidence). On-farm and off-farm livelihood diversification strategies increase the resilience of rural households against desertification and extreme weather events, such as droughts (high confidence). Strengthening collective action is important for addressing causes and impacts of desertification, and for adapting to climate change (medium confidence). A greater emphasis on understanding gender-specific differences over land use and land management practices can help make land restoration projects more successful (medium confidence). Improved access to markets raises agricultural profitability and motivates investment into climate change adaptation and SLM (medium confidence). Payments for ecosystem services give additional incentives to land users to adopt SLM practices (medium confidence). Expanding access to rural advisory services increases the knowledge on SLM and facilitates their wider adoption (medium confidence). Developing, enabling and promoting access to cleaner energy sources and technologies can contribute to reducing desertification and mitigating climate change through decreasing the use of fuelwood and crop residues for energy (medium confidence). Policy responses to droughts based on proactive drought preparedness and drought risk mitigation are more efficient in limiting drought-caused damages than reactive drought relief efforts (high confidence). {3.4.2, 3.6.2, 3.6.3, Cross-Chapter Box 5 in this chapter}

The knowledge on limits of adaptation to the combined effects of climate change and desertification is insufficient. However, the potential for residual risks and maladaptive outcomes is high (high confidence). Empirical evidence on the limits to adaptation in dryland areas is limited. Potential limits to adaptation include losses of land productivity due to irreversible forms of desertification. Residual risks can emerge from the inability of SLM measures to fully compensate for yield losses due to climate change impacts. They also arise from foregone reductions in ecosystem services due to soil fertility loss even when applying SLM measures could revert land to initial productivity after some time. Some activities favouring agricultural intensification in dryland areas can become maladaptive due to their negative impacts on the environment (medium confidence) Even when solutions are available, social, economic and institutional constraints could pose barriers to their implementation (medium confidence). {3.6.4}

Improving capacities, providing higher access to climate services, including local-level early warning systems, and expanding the use of remote sensing technologies are high-return investments for enabling effective adaptation and mitigation responses that help address desertification (high confidence). Reliable and timely climate services, relevant to desertification, can aid the development of appropriate adaptation and mitigation options reducing, the impact of desertification on human and natural systems (high confidence), with quantitative estimates showing that every USD invested in strengthening hydro-meteorological and early warning services in developing countries can yield between 4 and 35 USD (low confidence). Knowledge and flow of knowledge on desertification is currently fragmented. Improved knowledge and data exchange and sharing will increase the effectiveness of efforts to achieve Land Degradation Neutrality (high confidence). Expanded use of remotely sensed information for data collection helps in measuring progress towards achieving Land Degradation Neutrality (low evidence, high agreement). {3.2.1, 3.6.2, 3.6.3, Cross-Chapter Box 5 in this chapter}

3.1

The nature of desertification

3.1.1

Introduction

In this report, desertification is defined as land degradation in arid, semi-arid, and dry sub-humid areas resulting from many factors, including climatic variations and human activities (United Nations Convention to Combat Desertification (UNCCD) 1994). Land degradation is a negative trend in land condition, caused by direct or indirect human-induced processes including anthropogenic climate change, expressed as long-term reduction or loss of at least one of the following: biological productivity, ecological integrity or value to humans (Section 4.1.3). Arid, semi-arid, and dry sub-humid areas, together with hyper-arid areas, constitute drylands (UNEP 19921), home to about 3 billion people (van der Esch et al. 20172). The difference between desertification and land degradation is not process-based but geographic. Although land degradation can occur anywhere across the world, when it occurs in drylands, it is considered desertification (FAQ 1.3). Desertification is not limited to irreversible forms of land degradation, nor is it equated to desert expansion, but represents all forms and levels of land degradation occurring in drylands.

The geographic classification of drylands is often based on the aridity index (AI) – the ratio of average annual precipitation amount (P) to potential evapotranspiration amount (PET, see Glossary) (Figure 3.1). Recent estimates, based on AI, suggest that drylands cover about 46.2% (±0.8%) of the global land area (Koutroulis 20193; Prăvălie 20164) (low confidence). Hyper-arid areas, where the aridity index is below 0.05, are included in drylands, but are excluded from the definition of desertification (UNCCD 19945). Deserts are valuable ecosystems (UNEP 20066; Safriel 20097) geographically located in drylands and vulnerable to climate change. However, they are not considered prone to desertification. Aridity is a long-term climatic feature characterised by low average precipitation or available water (Gbeckor-Kove 19898; Türkeş 19999). Thus, aridity is different from drought, which is a temporary climatic event (Maliva and Missimer 201210). Moreover, droughts are not restricted to drylands, but occur both in drylands and humid areas (Wilhite et al. 201411). Following the Synthesis Report (SYR) of the IPCC Fifth Assessment Report (AR5), drought is defined here as “a period of abnormally dry weather long enough to cause a serious hydrological imbalance” (Mach et al. 201412) (Cross-Chapter Box 5 in this chapter).

AI is not an accurate proxy for delineating drylands in an increasing CO2 environment (Section 3.2.1). The suggestion that most of the world has become more arid, since the AI has decreased, is not supported by changes observed in precipitation, evaporation or drought (Sheffield et al. 201213; Greve et al. 201414). While climate change is expected to decrease the AI due to increases in potential evaporation, the assumptions that underpin the potential evaporation calculation are not consistent with a changing CO2 environment and the effect this has on transpiration rates (Roderick et al. 201515; Milly and Dunne 201616; Greve et al. 201717) (Section 3.2.1). Given that future climate is characterised by significant increases in CO2, the usefulness of currently applied AI thresholds to estimate dryland areas is limited under climate change. If instead of the AI, other variables such as precipitation, soil moisture, and primary productivity are used to identify dryland areas, there is no clear indication that the extent of drylands will change overall under climate change (Roderick et al. 201518; Greve et al. 201719; Lemordant et al. 201820). Thus, some dryland borders will expand, while some others will contract (high confidence).

Approximately 70% of dryland areas are located in Africa and Asia (Figure 3.2). The biggest land use/cover in terms of area in drylands, if deserts are excluded, are grasslands, followed by forests and croplands (Figure 3.3). The category of ‘other lands’ in Figure 3.3 includes bare soil, ice, rock, and all other land areas that are not included within the other five categories (FAO 201621). Thus, hyper-arid areas contain mostly deserts, with some small exceptions, for example, where grasslands and croplands are cultivated under oasis conditions with irrigation (Section 3.7.4). Moreover, FAO (2016)1786 defines grasslands as permanent pastures and meadows used continuously for more than five years. In drylands, transhumance, i.e. seasonal migratory grazing, often leads to non-permanent pasture systems, thus some of the areas under the ‘other land’ category are also used as non-permanent pastures (Ramankutty et al. 200822; Fetzel et al. 201723; Erb et al. 201624).

Figure 3.1

Geographical distribution of drylands, delimited based on the aridity index (AI). The classification of AI is: Humid AI > 0.65, Dry sub-humid 0.50 < AI ≤ 0.65, Semi-arid 0.20 < AI ≤ 0.50, Arid 0.05 < AI ≤ 0.20, Hyper-arid AI < 0.05. Data: TerraClimate precipitation and potential evapotranspiration (1980–2015) (Abatzoglou et al. 2018).

Geographical distribution of drylands, delimited based on the aridity index (AI). The classification of AI is: Humid AI > 0.65, Dry sub-humid 0.50 < AI ≤ 0.65, Semi-arid 0.20 < AI ≤ 0.50, Arid 0.05 < AI ≤ 0.20, Hyper-arid AI < 0.05. Data: TerraClimate precipitation and potential evapotranspiration (1980–2015) (Abatzoglou et al. 20181787).

Figure 3.2

Dryland categories across geographical areas (continents and Pacific region). Data: TerraClimate precipitation and potential evapotranspiration (1980–2015) (Abatzoglou et al. 2018).

Dryland categories across geographical areas (continents and Pacific region). Data: TerraClimate precipitation and potential evapotranspiration (1980–2015) (Abatzoglou et al. 20181788).

In the earlier global assessments of desertification (since the 1970s), which were based on qualitative expert evaluations, the extent of desertification was found to range between 4% and 70% of the area of drylands (Safriel 200725). More recent estimates, based on remotely sensed data, show that about 24–29% of the global land area experienced reductions in biomass productivity between the 1980s and 2000s (Bai et al. 200826; Le et al. 201627), corresponding to about 9.2% of drylands (±0.5%) experiencing declines in biomass productivity during this period (low confidence), mainly due to anthropogenic causes. Both of these studies consider rainfall dynamics, thus, accounting for the effect of droughts. While less than 10% of drylands is undergoing desertification, it is occurring in areas that contain around 20% of dryland population (Klein Goldewijk et al. 201728). In these areas the population has increased from approximately 172 million in 1950 to over 630 million today (Figure 1.1).

Available assessments of the global extent and severity of desertification are relatively crude approximations with considerable uncertainties, for example, due to confounding effects of invasive bush encroachment in some dryland regions. Different indicator sets and approaches have been developed for monitoring and assessment of desertification from national to global scales (Imeson 201229; Sommer et al. 201130; Zucca et al. 201231; Bestelmeyer et al. 201332). Many indicators of desertification only include a single factor or characteristic of desertification, such as the patch size distribution of vegetation (Maestre and Escudero 200933; Kéfi et al. 201034), Normalized Difference Vegetation Index (NDVI) (Piao et al. 200535), drought-tolerant plant species (An et al. 2007), grass cover (Bestelmeyer et al. 201336), land productivity dynamics (Baskan et al. 201737), ecosystem net primary productivity (Zhou et al. 201538) or Environmentally Sensitive Land Area Index (Symeonakis et al. 201639). In addition, some synthetic indicators of desertification have also been used to assess desertification extent and desertification processes, such as climate, land use, soil, and socio-economic parameters (Dharumarajan et al. 201840), or changes in climate, land use, vegetation cover, soil properties and population as the desertification vulnerability index (Salvati et al. 200941). Current data availability and methodological challenges do not allow for accurately and comprehensively mapping desertification at a global scale (Cherlet et al. 201842). However, the emerging partial evidence points to a lower global extent of desertification than previously estimated (medium confidence) (Section 3.2).

This assessment examines the socio-ecological links between drivers (Section 3.1) and feedbacks (Section 3.3) that influence desertification–climate change interactions, and then examines associated observed and projected impacts (Sections 3.4 and 3.5) and responses (Section 3.6). Moreover, this assessment highlights that dryland populations are highly vulnerable to desertification and climate change (Sections 3.2 and 3.4). At the same time, dryland populations also have significant past experience and sources of resilience embodied in indigenous and local knowledge and practices in order to successfully adapt to climatic changes and address desertification (Section 3.6). Numerous site-specific technological response options are also available for SLM in drylands that can help increase the resilience of agricultural livelihood systems to climate change (Section 3.6). However, continuing environmental degradation combined with climate change is straining the resilience of dryland populations. Enabling policy responses for SLM and livelihoods diversification can help maintain and strengthen the resilience and adaptive capacities in dryland areas (Section 3.6). The assessment finds that policies promoting SLM in drylands will contribute to climate change adaptation and mitigation, with co-benefits for broader sustainable development (high confidence) (Section 3.4).

Figure 3.3

Land use and land cover in drylands and share of each dryland category in global land area. Source: FAO (2016).

Land use and land cover in drylands and share of each dryland category in global land area. Source: FAO (2016)1789.

3.1.3

Dryland populations: Vulnerability and resilience

Drylands are home to approximately 38.2% (±0.6%) of the global population (Koutroulis 201951; van der Esch et al. 201752), that is about 3 billion people. The highest number of people live in the drylands of South Asia (Figure 3.4), followed by Sub-Saharan Africa and Latin America (van der Esch et al. 201753). In terms of the number of people affected by desertification, Reynolds et al. (2007) indicated that desertification was directly affecting 250 million people. More recent estimates show that 500 (±120) million people lived in 2015 in those dryland areas which experienced significant loss in biomass productivity between the 1980s and 2000s (Bai et al. 200854; Le et al. 201655). The highest numbers of affected people were in South and East Asia, North Africa and the Middle East (low confidence). The population in drylands is projected to increase about twice as rapidly as non-drylands, reaching 4 billion people by 2050 (van der Esch et al. 201756). This is due to higher population growth rates in drylands. About 90% of the population in drylands live in developing countries (UN-EMG 201157).

Figure 3.4

Current and projected population (under SSP2) in drylands, in billions. Source: van der Esch et al. (2017).

Current and projected population (under SSP2) in drylands, in billions. Source: van der Esch et al. (2017)1790.

Dryland populations are highly vulnerable to desertification and climate change because their livelihoods are predominantly dependent on agriculture, one of the sectors most susceptible to climate change (Rosenzweig et al. 201458; Schlenker and Lobell 201059). Climate change is projected to have substantial impacts on all types of agricultural livelihood systems in drylands (CGIAR-RPDS 201460) (Sections 3.4.1 and 3.4.2).

One key vulnerable group in drylands are pastoral and agropastoral households1. There are no precise figures about the number of people practicing pastoralism globally. Most estimates range between 100 million and 200 million (Rass 200661; Secretariat of the Convention on Biological Diversity 201062), of whom 30–63 million are nomadic pastoralists (Dong 201663; Carr-Hill 201364)2

Pastoral production systems represent an adaptation to high seasonal climate variability and low biomass productivity in dryland ecosystems (Varghese and Singh 201665; Krätli and Schareika 201066), which require large areas for livestock grazing through migratory pastoralism (Snorek et al. 201467). Grazing lands across dryland environments are being degraded, and/or being converted to crop production, limiting the opportunities for migratory livestock systems, and leading to conflicts with sedentary crop producers (Abbass 201468; Dimelu et al. 201669). These processes, coupled with ethnic differences, perceived security threats, and misunderstanding of pastoral rationality, have led to increasing marginalisation of pastoral communities and disruption of their economic and cultural structures (Elhadary 201470; Morton 201071). As a result, pastoral communities are not well prepared to deal with increasing weather/climate variability and weather/climate extremes due to changing climate (Dong 201672; López-i-Gelats et al. 201673), and remain amongst the most food insecure groups in the world (FAO 2018).

There is an increasing concentration of poverty in the dryland areas of Sub-Saharan Africa and South Asia (von Braun and Gatzweiler 201474; Barbier and Hochard 2016)75, where 41% and 12% of the total populations live in extreme poverty, respectively (World Bank 201876). For comparison, the average share of global population living in extreme poverty is about 10% (World Bank 201877). Multidimensional poverty, prevalent in many dryland areas, is a key source of vulnerability (Safriel et al. 200578; Thornton et al. 201479; Fraser et al. 201180; Thomas 200881). Multidimensional poverty incorporates both income-based poverty, and also other dimensions such as poor healthcare services, lack of education, lack of access to water, sanitation and energy, disempowerment, and threat from violence (Bourguignon and Chakravarty 200382; Alkire and Santos 201083, 201484). Contributing elements to this multidimensional poverty in drylands are rapid population growth, fragile institutional environment, lack of infrastructure, geographic isolation and low market access, insecure land tenure systems, and low agricultural productivity (Sietz et al. 201185; Reynolds et al. 201186; Safriel and Adeel 200887; Stafford Smith 201688). Even in high-income countries, those dryland areas that depend on agricultural livelihoods represent relatively poorer locations nationally, with fewer livelihood opportunities, for example in Italy (Salvati 201489). Moreover, in many drylands areas, female-headed households, women and subsistence farmers (both male and female) are more vulnerable to the impacts of desertification and climate change (Nyantakyi-Frimpong and Bezner-Kerr 201590; Sultana 201491; Rahman 201392). Some local cultural traditions and patriarchal relationships were found to contribute to higher vulnerability of women and female-headed households through restrictions on their access to productive resources (Nyantakyi-Frimpong and Bezner-Kerr 201594; Sultana 201495; Rahman 20131791) (Sections 3.4.2 and 3.6.3, and Cross-Chapter Box 11 in Chapter 7).

Despite these environmental, socio-economic and institutional constraints, dryland populations have historically demonstrated remarkable resilience, ingenuity and innovations, distilled into ILK to cope with high climatic variability and sustain livelihoods (Safriel and Adeel 200896; Davis 201697; Davies 201798) (Sections 3.6.1 and 3.6.2, and Cross-Chapter Box 13 in Chapter 7). For example, across the Arabian Peninsula and North Africa, informal community by-laws were successfully used for regulating grazing, collection and cutting of herbs and wood, and which limited rangeland degradation (Gari 200699; Hussein 2011100). Pastoralists in Mongolia developed indigenous classifications of pasture resources which facilitated ecologically optimal grazing practices (Fernandez-Gimenez 2000101) (Section 3.6.2). Currently, however, indigenous and local knowledge and practices are increasingly lost or can no longer cope with growing demands for land-based resources (Dominguez 2014102; Fernández-Giménez and Fillat Estaque 2012103; Hussein 2011104; Kodirekkala 2017105; Moreno-Calles et al. 2012106) (Section 3.4.2). Unsustainable land management is increasing the risks from droughts, floods and dust storms (Sections 3.4.2 and 3.5). Policy actions promoting the adoption of SLM practices in dryland areas, based on both indigenous and local knowledge and modern science, and expanding alternative livelihood opportunities outside agriculture can contribute to climate change adaptation and mitigation, addressing desertification, with co-benefits for poverty eradication and food security (high confidence) (Cowie et al. 2018107; Liniger et al. 2017108; Safriel and Adeel 2008109; Stafford-Smith et al. 2017110).

3.1.4

Processes and drivers of desertification under climate change

3.1.4.1

Processes of desertification and their climatic drivers

Processes of desertification are mechanisms by which drylands are degraded. Desertification consists of both biological and non-biological processes. These processes are classified under broad categories of degradation of physical, chemical and biological properties of terrestrial ecosystems. The number of desertification processes is large and they are extensively covered elsewhere (IPBES 2018a111; Lal 2016112; Racine 2008113; UNCCD 2017114). Section 4.2.1 and Tables 4.1 and 4.2 in Chapter 4 highlight those which are particularly relevant for this assessment in terms of their links to climate change and land degradation, including desertification.

Drivers of desertification are factors which trigger desertification processes. Initial studies of desertification during the early-to-mid 20th century attributed it entirely to human activities. In one of the influential publications of that time, Lavauden (1927)115 stated that: “Desertification is purely artificial. It is only the act of the man…” However, such a uni-causal view of desertification was shown to be invalid (Geist et al. 2004116; Reynolds et al. 2007117) (Sections 3.1.4.2 and 3.1.4.3). Tables 4.1 and 4.2 in Chapter 4 summarise the drivers, linking them to the specific processes of desertification and land degradation under changing climate.

Erosion refers to removal of soil by the physical forces of water, wind, or often caused by farming activities such as tillage (Ginoux et al. 2012118). The global estimates of soil erosion differ significantly, depending on scale, study period and method used (García-Ruiz et al. 2015119), ranging from approximately 20 Gt yr–1 to more than 200 Gt yr–1 (Boix-Fayos et al. 2006120; FAO 2015121). There is a significant potential for climate change to increase soil erosion by water, particularly in those regions where precipitation volumes and intensity are projected to increase (Panthou et al. 2014122; Nearing et al. 2015123). On the other hand, while it is a dominant form of erosion in areas such as West Asia and the Arabian Peninsula (Prakash et al. 2015124; Klingmüller et al. 2016125), there is limited evidence concerning climate change impacts on wind erosion (Tables 4.1 and 4.2 in Chapter 4, and Section 3.5).

Saline and sodic soils (see Glossary) occur naturally in arid, semi-arid and dry sub-humid regions of the world. Climate change or hydrological change can cause soil salinisation by increasing the mineralised groundwater level. However, secondary salinisation occurs when the concentration of dissolved salts in water and soil is increased by anthropogenic processes, mainly through poorly managed irrigation schemes. The threat of soil and groundwater salinisation induced by sea level rise and seawater intrusion are amplified by climate change (Section 4.9.7).

Global warming is expected to accelerate soil organic carbon (SOC) turnover, since the decomposition of the soil organic matter by microbial activity begins with low soil water availability, but this moisture is insufficient for plant productivity (Austin et al. 2004126) (Section 3.4.1.1). SOC is also lost due to soil erosion (Lal 2009127); therefore, in some dryland areas leading to SOC decline (Sections 3.3.3 and 3.5.2) and the transfer of carbon (C) from soil to the atmosphere (Lal 2009128).

Sea surface temperature (SST) anomalies can drive rainfall changes, with implications for desertification processes. North Atlantic SST anomalies are positively correlated with Sahel rainfall anomalies (Knight et al. 2006129; Gonzalez-Martin et al. 2014130; Sheen et al. 2017131). While the eastern tropical Pacific SST anomalies have a negative correlation with Sahel rainfall (Pomposi et al. 2016132), a cooler North Atlantic is related to a drier Sahel, with this relationship enhanced if there is a simultaneous relative warming of the South Atlantic (Hoerling et al. 2006133). Huber and Fensholt (2011)134 explored the relationship between SST anomalies and satellite observed Sahel vegetation dynamics, finding similar relationships but with substantial west–east variations in both the significant SST regions and the vegetation response. Concerning the paleoclimatic evidence on aridification after the early Holocene ‘Green Sahara’ period (11,000 to 5000 years ago), Tierney et al. (2017)135 indicate that a cooling of the North Atlantic played a role (Collins et al. 2017136; Otto-Bliesner et al. 2014137; Niedermeyer et al. 2009138) similar to that found in modern observations. Besides these SST relationships, aerosols have also been suggested as a potential driver of the Sahel droughts (Rotstayn and Lohmann 2002139; Booth et al. 2012140; Ackerley et al. 2011141). For eastern Africa, both recent droughts and decadal declines have been linked to human-induced warming in the western Pacific (Funk et al. 2018142).

Invasive plants contributed to desertification and loss of ecosystem services in many dryland areas in the last century (high confidence) (Section 3.7.3). Extensive woody plant encroachment altered runoff and soil erosion across much of the drylands, because the bare soil between shrubs is very susceptible to water erosion, mainly in high-intensity rainfall events (Manjoro et al. 2012143; Pierson et al. 2013144; Eldridge et al. 2015145). Rising CO2 levels due to global warming favour more rapid expansion of some invasive plant species in some regions. An example is the Great Basin region in western North America where over 20% of ecosystems have been significantly altered by invasive plants, especially exotic annual grasses and invasive conifers, resulting in loss of biodiversity. This land-cover conversion has resulted in reductions in forage availability, wildlife habitat, and biodiversity (Pierson et al. 2011, 2013146; Miller et al. 2013147).

The wildfire is a driver of desertification, because it reduces vegetation cover, increases runoff and soil erosion, reduces soil fertility and affects the soil microbial community (Vega et al. 2005148; Nyman et al. 2010149; Holden et al. 2013150; Pourreza et al. 2014151; Weber et al. 2014152; Liu and Wimberly 2016153). Predicted increases in temperature and the severity of drought events across some dryland areas (Section 2.2) can increase chances of wildfire occurrence (medium confidence) (Jolly et al. 2015154; Williams et al. 2010155; Clarke and Evans 2018156) (Cross-Chapter Box 3 in Chapter 2). In semi-arid and dry sub-humid areas, fire can have a profound influence on observed vegetation and particularly the relative abundance of grasses to woody plants (Bond et al. 2003157; Bond and Keeley 2005158; Balch et al. 2013159).

While large uncertainty exists concerning trends in droughts globally (AR5) (Section 2.2), examining the drought data by Ziese et al. (2014)160 for drylands only reveals a large inter-annual variability combined with a trend toward increasing dryland area affected by droughts since the 1950s (Figure 1.1). Thus, over the period 1961–2013, the annual area of drylands in drought has increased, on average, by slightly more than 1% per year, with large inter-annual variability.

3.1.4.2

Anthropogenic drivers of desertification under climate change

The literature on the human drivers of desertification is substantial (e.g., D’Odorico et al. 2013161; Sietz et al. 2011162; Yan and Cai 2015163; Sterk et al. 2016164; Varghese and Singh 2016165) and there have been several comprehensive reviews and assessments of these drivers very recently (Cherlet et al. 2018166; IPBES 2018a167; UNCCD 2017168). IPBES (2018a) identified cropland expansion, unsustainable land management practices including overgrazing by livestock, urban expansion, infrastructure development, and extractive industries as the main drivers of land degradation. IPBES (2018a) also found that the ultimate driver of land degradation is high and growing consumption of land-based resources, e.g., through deforestation and cropland expansion, escalated by population growth. What is particularly relevant in the context of the present assessment is to evaluate if, how and which human drivers of desertification will be modified by climate change effects.

Growing food demand is driving conversion of forests, rangelands, and woodlands into cropland (Bestelmeyer et al. 2015169; D’Odorico et al. 2013170). Climate change is projected to reduce crop yields across dryland areas (Sections 3.4.1 and 5.2.2), potentially reducing local production of food and feed. Without research breakthroughs mitigating these productivity losses through higher agricultural productivity, and reducing food waste and loss, meeting the increasing food demands of growing populations will require expansion of cropped areas to more marginal areas (with most prime areas in drylands already being under cultivation) (Lambin 2012171; Lambin et al. 2013172; Eitelberg et al. 2015173; Gutiérrez-Elorza 2006174; Kapović Solomun et al. 2018175). Borrelli et al. (2017)176 showed that the primary driver of soil erosion in 2012 was cropland expansion. Although local food demands could also be met by importing from other areas, this would mean increasing the pressure on land in those areas (Lambin and Meyfroidt 2011177). The net effects of such global agricultural production shifts on land condition in drylands are not known.

Climate change will exacerbate poverty among some categories of dryland populations (Sections 3.4.2 and 3.5.2). Depending on the context, this impact comes through declines in agricultural productivity, changes in agricultural prices and extreme weather events (Hertel and Lobell 2014178; Hallegatte and Rozenberg 2017179). There is high confidence that poverty limits both capacities to adapt to climate change and availability of financial resources to invest into SLM (Gerber et al. 2014180; Way 2016181; Vu et al. 2014182) (Sections 3.5.2, 3.6.2 and 3.6.3).

Labour mobility is another key human driver that will interact with climate change. Although strong impacts of climate change on migration in dryland areas are disputed, in some places, it is likely to provide an added incentive to migrate (Section 3.4.2.7). Out-migration will have several contradictory effects on desertification. On one hand, it reduces an immediate pressure on land if it leads to less dependence on land for livelihoods (Chen et al. 2014183; Liu et al. 2016a). Moreover, migrant remittances could be used to fund the adoption of SLM practices. Labour mobility from agriculture to non-agricultural sectors could allow land consolidation, gradually leading to mechanisation and agricultural intensification (Wang et al. 2014184, 2018185). On the other hand, this can increase the costs of labour-intensive SLM practices due to lower availability of rural agricultural labour and/or higher rural wages. Out-migration increases the pressure on land if higher wages that rural migrants earn in urban centres will lead to their higher food consumption. Moreover, migrant remittances could also be used to fund land-use expansion to marginal areas (Taylor et al. 2016186; Gray and Bilsborrow 2014187). The net effect of these opposite mechanisms varies from place to place (Qin and Liao 2016188). There is very little literature evaluating these joint effects of climate change, desertification and labour mobility (Section 7.3.2).

There are also many other institutional, policy and socio-economic drivers of desertification, such as land tenure insecurity, lack of property rights, lack of access to markets, and to rural advisory services, lack of technical knowledge and skills, agricultural price distortions, agricultural support and subsidies contributing to desertification, and lack of economic incentives for SLM (D’Odorico et al. 2013189; Geist et al. 2004190; Moussa et al. 2016191; Mythili and Goedecke 2016192; Sow et al. 2016193; Tun et al. 2015194; García-Ruiz 2010195). There is no evidence that these factors will be materially affected by climate change, however, serving as drivers of unsustainable land management practices, they do play a very important role in modulating responses for climate change adaptation and mitigation (Section 3.6.3).

3.1.4.3

Interaction of drivers: Desertification syndrome versus drylands development paradigm

Two broad narratives have historically emerged to describe responses of dryland populations to environmental degradation. The first is ‘desertification syndrome’ which describes the vicious cycle of resource degradation and poverty, whereby dryland populations apply unsustainable agricultural practices leading to desertification, and exacerbating their poverty, which then subsequently further limits their capacities to invest in SLM (MEA 2005196; Safriel and Adeel 2008197). The alternative paradigm is one of ‘drylands development’, which refers to social and technical ingenuity of dryland populations as a driver of dryland sustainability (MEA 2005; Reynolds et al. 2007198; Safriel and Adeel 2008199). The major difference between these two frameworks is that the ‘drylands development’ paradigm recognises that human activities are not the sole and/or most important drivers of desertification, but there are interactions of human and climatic drivers within coupled social-ecological systems (Reynolds et al. 2007200). This led Behnke and Mortimore (2016)201, and earlier Swift (1996)202, to conclude that the concept of desertification as irreversible degradation distorts policy and governance in dryland areas. Mortimore (2016)203 suggested that instead of externally imposed technical solutions, what is needed is for populations in dryland areas to adapt to this variable environment which they cannot control. All in all, there is high confidence that anthropogenic and climatic drivers interact in complex ways in causing desertification. As discussed in Section 3.2.2, the relative influence of human or climatic drivers on desertification varies from place to place (high confidence) (Bestelmeyer et al. 2018204; D’Odorico et al. 2013205; Geist and Lambin 2004206; Kok et al. 2016207; Polley et al. 2013208; Ravi et al. 2010209; Scholes 2009210; Sietz et al. 2017211; Sietz et al. 2011212).

3.2

Observations of desertification

3.2.2

Attribution of desertification

Desertification is a result of complex interactions within coupled social-ecological systems. Thus, the relative contributions of climatic, anthropogenic and other drivers of desertification vary depending on specific socio-economic and ecological contexts. The high natural climate variability in dryland regions is a major cause of vegetation changes but does not necessarily imply degradation. Drought is not degradation as the land productivity may return entirely once the drought ends (Kassas 1995410). However, if droughts increase in frequency, intensity and/or duration they may overwhelm the vegetation’s ability to recover (ecosystem resilience, Prince et al. 2018), causing degradation. Assuming a stationary climate and no human influence, rainfall variability results in fluctuations in vegetation dynamics which can be considered temporary, as the ecosystem tends to recover with rainfall, and desertification does not occur (Ellis 1995411; Vetter 2005412; von Wehrden et al. 2012413). Climate change on the other hand, exemplified by a non-stationary climate, can gradually cause a persistent change in the ecosystem through aridification and CO2 changes. Assuming no human influence, this ‘natural’ climatic version of desertification may take place rapidly, especially when thresholds are reached (Prince et al. 2018414), or over longer periods of time as the ecosystems slowly adjust to a new climatic norm through progressive changes in the plant community composition. Accounting for this climatic variability is required before attributions to other causes of desertification can be made.

For attributing vegetation changes to climate versus other causes, rain use efficiency (RUE – the change in net primary productivity (NPP) per unit of precipitation) and its variations in time have been used (Prince et al. 1998415). Global applications of RUE trends to attribute degradation to climate or other (largely human) causes have been performed by Bai et al. (2008)416 and Le et al. (2016)417 (Section 3.2.1.1). The RESTREND (residual trend) method analyses the correlation between annual maximum NDVI (or other vegetation index as a proxy for NPP) and precipitation by testing accumulation and lag periods for the precipitation (Evans and Geerken 2004418). The identified relationship with the highest correlation represents the maximum amount of vegetation variability that can be explained by the precipitation, and corresponding RUE values can be calculated. Using this relationship, the climate component of the NDVI time series can be reconstructed, and the difference between this and the original time series (the residual) is attributed to anthropogenic and other causes.

The RESTREND method, or minor variations of it, have been applied extensively. Herrmann and Hutchinson (2005)419 concluded that climate was the dominant causative factor for widespread greening in the Sahel region from 1982–2003, and anthropogenic and other factors were mostly producing land improvements or no change. However, pockets of desertification were identified in Nigeria and Sudan. Similar results were also found from 1982–2007 by Huber et al. (2011)420. Wessels et al. (2007)421 applied RESTREND to South Africa and showed that RESTREND produced a more accurate identification of degraded land than RUE alone. RESTREND identified a smaller area undergoing desertification due to non-climate causes compared to the NDVI trends. Liu et al. (2013)430 extended the climate component of RESTREND to include temperature and applied this to VOD observations of the cold drylands of Mongolia. They found the area undergoing desertification due to non-climatic causes is much smaller than the area with negative VOD trends. RESTREND has also been applied in several other studies to the Sahel (Leroux et al. 2017422), Somalia (Omuto et al. 2010)423, West Africa (Ibrahim et al. 2015)424, China (Li et al. 2012425; Yin et al. 2014426), Central Asia (Jiang et al. 2017427), Australia (Burrell et al. 2017428) and globally (Andela et al. 2013429). In each of these studies the extent to which desertification can be attributed to climate versus other causes varies across the landscape.

These studies represent the best regional, remote-sensing based attribution studies to date, noting that RESTREND and RUE have some limitations (Higginbottom and Symeonakis 2014431). Vegetation growth (NPP) changes slowly compared to rainfall variations and may be sensitive to rainfall over extended periods (years), depending on vegetation type. Detection of lags and the use of weighted antecedent rainfall can partially address this problem, though most studies do not do this. The method addresses changes since the start of the time series; it cannot identify whether an area is already degraded at the start time. It is assumed that climate, particularly rainfall, is a principal factor in vegetation change which may not be true in more humid regions.

Another assumption in RESTREND is that any trend is linear throughout the period examined. That is, there are no discontinuities (break points) in the trend. Browning et al. (2017)432 have shown that break points in NDVI time series reflect vegetation changes based on long-term field sites. To overcome this limitation, Burrell et al. (2017)433 introduced the Time Series Segmentation-RESTREND (TSS-RESTREND) which allows a breakpoint or turning point within the period examined (Figure 3.7). Using TSS-RESTREND over Australia they identified more than double the degrading area than could be identified with a standard RESTREND analysis. The occurrence and drivers of abrupt change (turning points) in ecosystem functioning were also examined by Horion et al. (2016)434 over the semi-arid Northern Eurasian agricultural frontier. They combined trend shifts in RUE, field data and expert knowledge, to map environmental hotspots of change and attribute them to climate and human activities. One-third of the area showed significant change in RUE, mainly occurring around the fall of the Soviet Union (1991) or as the result of major droughts. Recent human-induced turning points in ecosystem functioning were uncovered around Volgograd (Russia) and around Lake Balkhash (Kazakhstan), attributed to recultivation, increased salinisation, and increased grazing.

Attribution of vegetation changes to human activity has also been done within modelling frameworks. In these methods ecosystem models are used to simulate potential natural vegetation dynamics, and this is compared to the observed state. The difference is attributed to human activities. Applied to the Sahel region during the period of 1982–2002, it showed that people had a minor influence on vegetation changes (Seaquist et al. 2009435). Similar model/observation comparisons performed globally found that CO2 fertilisation was the strongest forcing at global scales, with climate having regionally varying effects (Mao et al. 2013436; Zhu et al. 2016437). Land-use/ land-cover change was a dominant forcing in localised areas. The use of this method to examine vegetation changes in China (1982–2009) attributed most of the greening trend to CO2 fertilisation and nitrogen (N) deposition (Piao et al. 2015). However in some parts of northern and western China, which includes large areas of drylands, Piao et al. (2015)438 found climate changes could be the dominant forcing. In the northern extratropical land surface, the observed greening was consistent with increases in greenhouse gases (notably CO2) and the related climate change, and not consistent with a natural climate that does not include anthropogenic increase in greenhouse gases (Mao et al. 2016439). While many studies found widespread influence of CO2 fertilisation, it is not ubiquitous; for example, Lévesque et al. (2014) found little response to CO2 fertilisation in some tree species in Switzerland/northern Italy.

Using multiple extreme-event attribution methodologies, Uhe et al. (2018)440 shows that the dominant influence for droughts in eastern Africa during the October–December ‘short rains’ season is the prevailing tropical SST patterns, although temperature trends mean that the current drought conditions are hotter than they would have been without climate change. Similarly, Funk et al. (2019)441 found that 2017 March–June East African drought was influenced by Western Pacific SST, with high SST conditions attributed to climate change.

There are numerous local case studies on attribution of desertification, which use different periods, focus on different land uses and covers, and consider different desertification processes. For example, two-thirds of the observed expansion of the Sahara Desert from 1920–2003 has been attributed to natural climate cycles (the cold phase of Atlantic Multi-Decadal Oscillation and Pacific Decadal Oscillation) (Thomas and Nigam 2018442). Some studies consider drought to be the main driver of desertification in Africa (e.g., Masih et al. 2014443). However, other studies suggest that although droughts may contribute to desertification, the underlying causes are human activities (Kouba et al. 2018444). Brandt et al. (2016a) found that woody vegetation trends are negatively correlated with human population density. Changes in land use, water pumping and flow diversion have enhanced drying of wetlands and salinisation of freshwater aquifers in Israel (Inbar 2007445). The dryland territory of China has been found to be very sensitive to both climatic variations and land-use/land-cover changes (Fu et al. 2000446; Liu and Tian 2010447; Zhao et al. 2013, 2006448). Feng et al. (2015) shows that socio-economic factors were dominant in causing desertification in north Shanxi, China, between 1983 and 2012, accounting for about 80% of desertification expansion. Successful grass establishment has been impeded by overgrazing and nutrient depletion leading to the encroachment of shrubs into the northern Chihuahuan Desert (USA) since the mid-19th century (Kidron and Gutschick 2017449). Human activities led to rangeland degradation in Pakistan and Mongolia during 2000–2011 (Lei et al. 2011450). More equal shares of climatic (temperature and precipitation trends) and human factors were attributed for changes in rangeland condition in China (Yang et al. 2016451).

This kaleidoscope of local case studies demonstrates how attribution of desertification is still challenging for several reasons. Firstly, desertification is caused by an interaction of different drivers which vary in space and time. Secondly, in drylands, vegetation reacts closely to changes in rainfall so the effect of rainfall changes on biomass needs to be ‘removed’ before attributing desertification to human activities. Thirdly, human activities and climatic drivers impact vegetation/ ecosystem changes at different rates. Finally, desertification manifests as a gradual change in ecosystem composition and structure (e.g., woody shrub invasion into grasslands). Although initiated at a limited location, ecosystem change may propagate throughout an extensive area via a series of feedback mechanisms. This complicates the attribution of desertification to human and climatic causes, as the process can develop independently once started.

Figure 3.7

The drivers of dryland vegetation change. The mean annual change in NDVImax between 1982 and 2015 (see Figure 3.6 for total change using Global Inventory Modelling and Mapping Studies NDVI3g v1 dataset) attributable to(a)CO2 fertilisation(b)climate and (c) land use. The change attributable to CO2 fertilisation was calculated using the CO2 fertilisation relationship described in Franks […]

The drivers of dryland vegetation change. The mean annual change in NDVImax between 1982 and 2015 (see Figure 3.6 for total change using Global Inventory Modelling and Mapping Studies NDVI3g v1 dataset) attributable to(a)CO2 fertilisation(b)climate and (c) land use. The change attributable to CO2 fertilisation was calculated using the CO2 fertilisation relationship described in Franks et al. 20131793. The Time Series Segmented Residual Trends (TSS-RESTREND) method (Burrell et al. 20171794) applied to the CO2-adjusted NDVI was used to separate Climate and Land Use. A multi-climate dataset ensemble was used to reduce the impact of dataset errors (Burrell et al. 20181795). Non-dryland regions (aridity index >0.65) are masked in dark grey. Areas where the change did not meet the multi-run ensemble significance criteria, or are smaller than the error in the sensors (±0.00001) are masked in white

Rasmussen et al. (2016)452 studied the reasons behind the overall lack of scientific agreement in trends of environmental changes in the Sahel, including their causes. The study indicated that these are due to differences in conceptualisations and choice of indicators, biases in study site selection, differences in methods, varying measurement accuracy, differences in time and spatial scales. High-resolution, multi-sensor airborne platforms provide a way to address some of these issues (Asner et al. 2012453).

The major conclusion of this section is that, with all the shortcomings of individual case studies, relative roles of climatic and human drivers of desertification are location-specific and evolve over time (high confidence). Biophysical research on attribution and socio-economic research on drivers of land degradation have long studied the same topic, but in parallel, with little interdisciplinary integration. Interdisciplinary work to identify typical patterns, or typologies, of such interactions of biophysical and human drivers of desertification (not only of dryland vulnerability), and their relative shares, done globally in comparable ways, will help in the formulation of better informed policies to address desertification and achieve land degradation neutrality.

3.3

Desertification feedbacks to climate

While climate change can drive desertification (Section 3.1.4.1), the process of desertification can also alter the local climate, providing a feedback (Sivakumar 2007454). These feedbacks can alter the carbon cycle, and hence the level of atmospheric CO2 and its related global climate change, or they can alter the surface energy and water budgets, directly impacting the local climate. While these feedbacks occur in all climate zones (Chapter 2), here we focus on their effects in dryland regions and assess the literature concerning the major desertification feedbacks to climate. The main feedback pathways discussed throughout Section 3.3 are summarised in Figure 3.8.

Drylands are characterised by limited soil moisture compared to humid regions. Thus, the sensible heat (heat that causes the atmospheric temperature to rise) accounts for more of the surface net radiation than latent heat (evaporation) in these regions (Wang and Dickinson 2013455). This tight coupling between the surface energy balance and the soil moisture in semi-arid and dry sub-humid zones makes these regions susceptible to land–atmosphere feedback loops that can amplify changes to the water cycle (Seneviratne et al. 2010456). Changes to the land surface caused by desertification can change the surface energy budget, altering the soil moisture and triggering these feedbacks.

Figure 3.8

Schematic of main pathways through which desertification can feed back on climate, as discussed in Section 3.4. Note: Red arrows indicate a positive effect. Blue arrows indicate a negative effect. Grey arrows indicate an indeterminate effect (potentially both positive and negative). Solid arrows are direct while dashed arrows are indirect.

Schematic of main pathways through which desertification can feed back on climate, as discussed in Section 3.4. Note: Red arrows indicate a positive effect. Blue arrows indicate a negative effect. Grey arrows indicate an indeterminate effect (potentially both positive and negative). Solid arrows are direct while dashed arrows are indirect.

3.3.1

Sand and dust aerosols

Sand and mineral dust are frequently mobilised from sparsely vegetated drylands forming ‘sand storms’ or ‘dust storms’ (UNEP et al. 2016457). The African continent is the most important source of desert dust; perhaps 50% of atmospheric dust comes from the Sahara (Middleton 2017458). Ginoux et al. (2012)459 estimated that 25% of global dust emissions have anthropogenic origins, often in drylands. These events can play an important role in the local energy balance. Through reducing vegetation cover and drying the surface conditions, desertification can increase the frequency of these events. Biological or structural soil crusts have been shown to effectively stabilise dryland soils. Thus their loss due to intense land use and/ or climate change can be expected to cause an increase in sand and dust storms (high confidence) (Rajot et al. 2003460; Field et al. 2010461; Rodriguez-Caballero et al. 2018462). These sand and dust aerosols impact the regional climate in several ways (Choobari et al. 2014463). The direct effect is the interception, reflection and absorption of solar radiation in the atmosphere, reducing the energy available at the land surface and increasing the temperature of the atmosphere in layers with sand and dust present (Kaufman et al. 2002464; Middleton 2017465; Kok et al. 2018466). The heating of the dust layer can alter the relative humidity and atmospheric stability, which can change cloud lifetimes and water content. This has been referred to as the semi-direct effect (Huang et al. 2017467). Aerosols also have an indirect effect on climate through their role as cloud condensation nuclei, changing cloud radiative properties as well as the evolution and development of precipitation (Kaufman et al. 2002468). While these indirect effects are more variable than the direct effects, depending on the types and amounts of aerosols present, the general tendency is toward an increase in the number, but a reduction in the size of cloud droplets, increasing the cloud reflectivity and decreasing the chances of precipitation. These effects are referred to as aerosol-radiation and aerosol–cloud interactions (Boucher et al. 2013469).

There is high confidence that there is a negative relationship between vegetation green-up and the occurrence of dust storms (Engelstaedter et al. 2003470; Fan et al. 2015471; Yu et al. 2015472; Zou and Zhai 2004473). Changes in groundwater can affect vegetation and the generation of atmospheric dust (Elmore et al. 2008474). This can occur through groundwater processes such as the vertical movement of salt to the surface causing salinisation, supply of near-surface soil moisture, and sustenance of groundwater dependent vegetation. Groundwater dependent ecosystems have been identified in many dryland regions around the world (Decker et al. 2013475; Lamontagne et al. 2005476; Patten et al. 2008477). In these locations declining groundwater levels can decrease vegetation cover. Cook et al. (2009)478 found that dust aerosols intensified the ‘Dust Bowl’ drought in North America during the 1930s.

By decreasing the amount of green cover and hence increasing the occurrence of sand and dust storms, desertification will increase the amount of shortwave cooling associated with the direct effect (high confidence). There is medium confidence that the semi-direct and indirect effects of this dust would tend to decrease precipitation and hence provide a positive feedback to desertification (Huang et al. 2009479; Konare et al. 2008480; Rosenfeld et al. 2001481; Solmon et al. 2012482; Zhao et al. 2015483). However, the combined effect of dust has also been found to increase precipitation in some areas (Islam and Almazroui 2012484; Lau et al. 2009485; Sun et al. 2012486). The overall combined effect of dust aerosols on desertification remains uncertain with low agreement between studies that find positive (Huang et al. 2014487), negative (Miller et al. 2004488) or no feedback on desertification (Zhao et al. 2015489).

3.3.1.1

Off-site feedbacks

Aerosols can act as a vehicle for the long-range transport of nutrients to oceans (Jickells et al. 2005490; Okin et al. 2011491) and terrestrial land surfaces (Das et al. 2013492). In several locations, notably the Atlantic Ocean, the west of northern Africa, and the Pacific Ocean east of northern China, a considerable amount of mineral dust aerosols, sourced from nearby drylands, reaches the oceans. It was estimated that 60% of dust transported off Africa is deposited in the Atlantic Ocean (Kaufman et al. 2005493), while 50% of the dust generated in Asia reaches the Pacific Ocean or further (Uno et al. 2009494; Zhang et al. 1997495). The Sahara is also a major source of dust for the Mediterranean basin (Varga et al. 2014496). The direct effect of atmospheric dust over the ocean was found to be a cooling of the ocean surface (limited evidence, high agreement) (Evan and Mukhopadhyay 2010497; Evan et al. 2009498) with the tropical North Atlantic mixed layer cooling by over 1°C (Evan et al. 2009499).

It has been suggested that dust may act as a source of nutrients for the upper ocean biota, enhancing the biological activity and related carbon sink (medium evidence, low agreement) (Lenes et al. 2001500; Shaw et al. 2008501; Neuer et al. 2004502). The overall response depends on the environmental controls on the ocean biota, the type of aerosols including their chemical constituents, and the chemical environment in which they dissolve (Boyd et al. 2010503).

Dust deposited on snow can increase the amount of absorbed solar radiation leading to more rapid melting (Painter et al. 2018504), impacting a region’s hydrological cycle (high confidence). Dust deposition on snow and ice has been found in many regions of the globe (e.g., Painter et al. 2018; Kaspari et al. 2014505; Qian et al. 2015506; Painter et al. 2013507), however quantification of the effect globally and estimation of future changes in the extent of this effect remain knowledge gaps.

3.3.2

Changes in surface albedo

Increasing surface albedo in dryland regions will impact the local climate, decreasing surface temperature and precipitation, and provide a positive feedback on the albedo (high confidence) (Charney et al. 1975508). This albedo feedback can occur in desert regions worldwide (Zeng and Yoon 2009509). Similar albedo feedbacks have also been found in regional studies over the Middle East (Zaitchik et al. 2007510), Australia (Evans et al. 2017511; Meng et al. 2014a512, b513), South America (Lee and Berbery 2012514) and the USA (Zaitchik et al. 2013515).

Recent work has also found albedo in dryland regions can be associated with soil surface communities of lichens, mosses and cyanobacteria (Rodriguez-Caballero et al. 2018516). These communities compose the soil crust in these ecosystems and due to the sparse vegetation cover, directly influence the albedo. These communities are sensitive to climate changes, with field experiments indicating albedo changes greater than 30% are possible. Thus, changes in these communities could trigger surface albedo feedback processes (limited evidence, high agreement) (Rutherford et al. 2017517).

A further pertinent feedback relationship exists between changes in land-cover, albedo, carbon stocks and associated GHG emissions, particularly in drylands with low levels of cloud cover. One of the first studies to focus on the subject was Rotenberg and Yakir (2010)518, who used the concept of ‘radiative forcing’ to compare the relative climatic effect of a change in albedo with a change in atmospheric GHGs due to the presence of forest within drylands. Based on this analysis, it was estimated that the change in surface albedo due to the degradation of semi-arid areas has decreased radiative forcing in these areas by an amount equivalent to approximately 20% of global anthropogenic GHG emissions between 1970 and 2005.

3.3.3

Changes in vegetation and greenhouse gas fluxes

Terrestrial ecosystems have the ability to alter atmospheric GHGs through a number of processes (Schlesinger et al. 1990519). This may be through a change in plant and soil carbon stocks, either sequestering atmospheric CO2 during growth or releasing carbon during combustion and respiration, or through processes such as enteric fermentation of domestic and wild ruminants that leads to the release of methane and nitrous oxide (Sivakumar 2007520). It is estimated that 241–470 GtC is stored in dryland soils (top 1 m) (Lal 2004521; Plaza et al. 2018522). When evaluating the effect of desertification, the net balance of all the processes and associated GHG fluxes needs to be considered.

Desertification usually leads to a loss in productivity and a decline in above – and below-ground carbon stocks (Abril et al. 2005523; Asner et al. 2003524). Drivers such as overgrazing lead to a decrease in both plant and SOC pools (Abdalla et al. 2018525). While dryland ecosystems are often characterised by open vegetation, not all drylands have low biomass and carbon stocks in an intact state (Lechmere-Oertel et al. 2005526; Maestre et al. 2012527). Vegetation types such as the subtropical thicket of South Africa have over 70 tC ha–1 in an intact state, greater than 60% of which is released into the atmosphere during degradation through overgrazing (Lechmere-Oertel et al. 2005528; Powell 2009529). In comparison, semi-arid grasslands and savannahs with similar rainfall, may have only 5–35 tC ha–1 (Scholes and Walker 1993530; Woomer et al. 2004531).

At the same time, it is expected that a decline in plant productivity may lead to a decrease in fuel loads and a reduction in CO2, nitrous oxide and methane emissions from fire. In a similar manner, decreasing productivity may lead to a reduction in numbers of ruminant animals that in turn would decrease methane emissions. Few studies have focussed on changes in these sources of emissions due to desertification and it remains a field that requires further research.

In comparison to desertification through the suppression of primary production, the process of woody plant encroachment can result in significantly different climatic feedbacks. Increasing woody plant cover in open rangeland ecosystems leads to an increase in woody carbon stocks both above – and below-ground (Asner et al. 2003532; Hughes et al. 2006533; Petrie et al. 2015534; Li et al. 2016535). Within the drylands of Texas, USA, shrub encroachment led to a 32% increase in aboveground carbon stocks over a period of 69 years (3.8 tC ha–1 to 5.0 tC ha–1) (Asner et al. 2003536). Encroachment by taller woody species can lead to significantly higher observed biomass and carbon stocks. For example, encroachment by Dichrostachys cinerea and several Vachellia species in the sub-humid savannahs of north-west South Africa led to an increase of 31–46 tC ha–1 over a 50–65 year period (1936–2001) (Hudak et al. 2003537). In terms of potential changes in SOC stocks, the effect may be dependent on annual rainfall and soil type. Woody cover generally leads to an increase in SOC stocks in drylands that have less than 800 mm of annual rainfall, while encroachment can lead to a loss of soil carbon in more humid ecosystems (Barger et al. 2011538; Jackson et al. 2002539).

The suppression of the grass layer through the process of woody encroachment may lead to a decrease in carbon stocks within this relatively small carbon pool (Magandana 2016540). Conversely, increasing woody cover may lead to a decrease and even halt in surface fires and associated GHG emissions. In their analysis of drivers of fire in southern Africa, Archibald et al. (2009)541 note that there is a potential threshold around 40% canopy cover, above which surface grass fires are rare. Whilst there have been a number of studies on changes in carbon stocks due to desertification in North America, southern Africa and Australia, a global assessment of the net change in carbon stocks – as well as fire and ruminant GHG emissions due to woody plant encroachment – has not been done yet.

3.4

Desertification impacts on natural and socio-economic systems under climate change

3.4.1

Impacts on natural and managed ecosystems

3.4.1.1

Impacts on ecosystems and their services in drylands

The Millenium Ecosystem Assessement (2005)542 proposed four classes of ecosystem services: provisioning, regulating, supporting and cultural services (Cross-Chapter Box 8 in Chapter 6). These ecosystem services in drylands are vulnerable to the impacts of climate change due to high variability in temperature, precipitation and soil fertility (Enfors and Gordon 2008543; Mortimore 2005544). There is high confidence that desertification processes such as soil erosion, secondary salinisation, and overgrazing have negatively impacted provisioning ecosystem services in drylands, particularly food and fodder production (Majeed and Muhammad 2019545; Mirzabaev et al. 2016546; Qadir et al. 2009547; Van Loo et al. 2017548; Tokbergenova et al. 2018549) (Section 3.4.2.2). Zika and Erb (2009)550 reported an estimation of NPP losses between 0.8 and 2.0 GtC yr–1 due to desertification, comparing the potential NPP and the NPP calculated for the year 2000. In terms of climatic factors, although climatic changes between 1976 and 2016 were found to be favourable for crop yields overall in Russia (Ivanov et al. 2018551), yield decreases of up to 40–60% in dryland areas were caused by severe and extensive droughts (Ivanov et al. 2018552). Increase in temperature can have a direct impact on animals in the form of increased physiological stress (Rojas-Downing et al. 2017553), increased water requirements for drinking and cooling, a decrease in the production of milk, meat and eggs, increased stress during conception and reproduction (Nardone et al. 2010554) or an increase in seasonal diseases and epidemics (Thornton et al. 2009555; Nardone et al. 2010556). Furthermore, changes in temperature can indirectly impact livestock through reducing the productivity and quality of feed crops and forages (Thornton et al. 2009557; Polley et al. 2013558). On the other hand, fewer days with extreme cold temperatures during winter in the temperate zones are associated with lower livestock mortality. The future projection of impacts on ecosystems is presented in Section 3.5.2.

Over-extraction is leading to groundwater depletion in many dryland areas (high confidence) (Mudd 2000559; Mays 2013560; Mahmod and Watanabe 2014561; Jolly et al. 2008562). Globally, groundwater reserves have been reduced since 1900, with the highest rate of estimated reductions of 145 km3 yr–1 between 2000 and 2008 (Konikow 2011563). Some arid lands are very vulnerable to groundwater reductions, because the current natural recharge rates are lower than during the previous wetter periods (e.g., the Atacama Desert, and Nubian aquifer system in Africa) (Squeo et al. 2006564; Mahmod and Watanabe 2014565; Herrera et al. 2018566).

Among regulating services, desertification can influence levels of atmospheric CO2. In drylands, the majority of carbon is stored below ground in the form of biomass and SOC (FAO 1995567) (Section 3.3.3). Land-use changes often lead to reductions in SOC and organic matter inputs into soil (Albaladejo et al. 2013568; Almagro et al. 2010569; Hoffmann et al. 2012570; Lavee et al. 1998571; Rey et al. 2011572), increasing soil salinity and soil erosion (Lavee et al. 1998573; Martinez-Mena et al. 2008574). In addition to the loss of soil, erosion reduces soil nutrients and organic matter, thereby impacting land’s productive capacity. To illustrate, soil erosion by water is estimated to result in the loss of 23–42 Mt of nitrogen and 14.6–26.4 Mt of phosphorus from soils globally each year (Pierzynski et al. 2017575).

Precipitation, by affecting soil moisture content, is considered to be the principal determinant of the capacity of drylands to sequester carbon (Fay et al. 2008576; Hao et al. 2008577; Mi et al. 2015578; Serrano-Ortiz et al. 2015579; Vargas et al. 2012580; Sharkhuu et al. 2016581). Lower annual rainfall resulted in the release of carbon into the atmosphere for a number of sites located in Mongolia, China and North America (Biederman et al. 2017582; Chen et al. 2009583; Fay et al. 2008584; Hao et al. 2008585; Mi et al. 2015586; Sharkhuu et al. 2016587). Low soil water availability promotes soil microbial respiration, yet there is insufficient moisture to stimulate plant productivity (Austin et al. 2004588), resulting in net carbon emissions at an ecosystem level. Under even drier conditions, photo degradation of vegetation biomass may often constitute an additional loss of carbon from an ecosystem (Rutledge et al. 2010589). In contrast, years of good rainfall in drylands resulted in the sequestration of carbon (Biederman et al. 2017590; Chen et al. 2009591; Hao et al. 2008592). In an exceptionally rainy year (2011) in the southern hemisphere, the semi-arid ecosystems of this region contributed 51% of the global net carbon sink (Poulter et al. 2014593). These results suggest that arid ecosystems could be an important global carbon sink, depending on soil water availability (medium evidence, high agreement). However, drylands are generally predicted to become warmer with an increasing frequency of extreme drought and high rainfall events (Donat et al. 2016594).

When desertification reduces vegetation cover, this alters the soil surface, affecting the albedo and the water balance (Gonzalez-Martin et al. 2014595) (Section 3.3). In such situations, erosive winds have no more obstacles, which favours the occurrence of wind erosion and dust storms. Mineral aerosols have an important influence on the dispersal of soil nutrients and lead to changes in soil characteristics (Goudie and Middleton 2001596; Middleton 2017597). Thereby, the soil formation as a supporting ecosystem service is negatively affected (Section 3.3.1). Soil erosion by wind results in a loss of fine soil particles (silt and clay), reducing the ability of soil to sequester carbon (Wiesmeier et al. 2015598). Moreover, dust storms reduce crop yields by loss of plant tissue caused by sandblasting (resulting in loss of plant leaves and hence reduced photosynthetic activity (Field et al. 2010599), exposing crop roots, crop seed burial under sand deposits, and leading to losses of nutrients and fertiliser from topsoil (Stefanski and Sivakumar 2009600)). Dust storms also impact crop yields by reducing the quantity of water available for irrigation; they can decrease the storage capacity of reservoirs by siltation, and block conveyance canals (Middleton 2017601; Middleton and Kang 2017602; Stefanski and Sivakumar 2009603). Livestock productivity is reduced by injuries caused by dust storms (Stefanski and Sivakumar 2009604). Additionally, dust storms favour the dispersion of microbial and plant species, which can make local endemic species vulnerable to extinction and promote the invasion of plant and microbial species (Asem and Roy 2010605; Womack et al. 2010606). Dust storms increase microbial species in remote sites (high confidence) (Kellogg et al. 2004607; Prospero et al. 2005608; Griffin et al. 2006609; Schlesinger et al. 2006610; Griffin 2007611; De Deckker et al. 2008612; Jeon et al. 2011613; Abed et al. 2012614; Favet et al. 2013615; Woo et al. 2013616; Pointing and Belnap 2014617).

3.4.1.2

Impacts on biodiversity: Plant and wildlife

Plant biodiversity

Over 20% of global plant biodiversity centres are located within drylands (White and Nackoney 2003618). Plant species located within these areas are characterised by high genetic diversity within populations (Martínez-Palacios et al. 1999619). The plant species within these ecosystems are often highly threatened by climate change and desertification (Millennium Ecosystem Assessment 2005b620; Maestre et al. 2012621). Increasing aridity exacerbates the risk of extinction of some plant species, especially those that are already threatened due to small populations or restricted habitats (Gitay et al. 2002622). Desertification, including through land-use change, already contributed to the loss of biodiversity across drylands (medium confidence) (Newbold et al. 2015623; Wilting et al. 2017624). For example, species richness decreased from 234 species in 1978 to 95 in 2011 following long periods of drought and human driven degradation on the steppe land of south-western Algeria (Observatoire du Sahara et du Sahel 2013625). Similarly, drought and overgrazing led to loss of biodiversity in Pakistan to the point that only drought-adapted species can now survive on the arid rangelands (Akhter and Arshad 2006626). Similar trends were observed in desert steppes of Mongolia (Khishigbayar et al. 2015627). In contrast, the increase in annual moistening of southern European Russia from the late 1980s to the beginning of the 21st century caused the restoration of steppe vegetation, even under conditions of strong anthropogenic pressure (Ivanov et al. 2018628). The seed banks of annual species can often survive over the long term, germinating in wet years, suggesting that these species could be resilient to some aspects of climate change (Vetter et al. 2005629). Yet, Hiernaux and Houérou (2006)630 showed that overgrazing in the Sahel tended to decrease the seed bank of annuals, which could make them vulnerable to climate change over time. Perennial species, considered as the structuring element of the ecosystem, are usually less affected as they have deeper roots, xeromorphic properties and physiological mechanisms that increase drought tolerance (Le Houérou 1996631). However, in North Africa, long-term monitoring (1978–2014) has shown that important plant perennial species have also disappeared due to drought (Stipa tenacissima and Artemisia herba alba) (Hirche et al. 2018633; Observatoire du Sahara et du Sahel 2013634). The aridisation of the climate in the south of Eastern Siberia led to the advance of the steppes to the north and to the corresponding migration of steppe mammal species between 1976 and 2016 (Ivanov et al. 2018635). The future projection of impacts on plant biodiversity is presented in Section 3.5.2.

Wildlife biodiversity

Dryland ecosystems have high levels of faunal diversity and endemism (MEA 2005636; Whitford 2002637). Over 30% of the endemic bird areas are located within these regions, which is also home to 25% of vertebrate species (Maestre et al. 2012638; MEA 2005639). Yet, many species within drylands are threatened with extinction (Durant et al. 2014640; Walther 2016641). Habitat degradation and desertification are generally associated with biodiversity loss (Ceballos et al. 2010642; Tang et al. 2018643; Newbold et al. 2015644). The ‘grazing value’ of land declines with both a reduction in vegetation cover and shrub encroachment, with the former being more detrimental to native vertebrates (Parsons et al. 2017645). Conversely, shrub encroachment may buffer desertification by increasing resource and microclimate availability, resulting in an increase in vertebrate species abundance and richness observed in the shrub-encroached arid grasslands of North America (Whitford 1997646) and Australia (Parsons et al. 2017647). However, compared to historically resilient drylands, these encroached habitats and their new species assemblages may be more sensitive to droughts, which may become more prevalent with climate change (Schooley et al. 2018648). Mammals and birds may be particularly sensitive to droughts because they rely on evaporative cooling to maintain their body temperatures within an optimal range (Hetem et al. 2016649) and risk lethal dehydration in water limited environments (Albright et al. 2017650). The direct effects of reduced rainfall and water availability are likely to be exacerbated by the indirect effects of desertification through a reduction in primary productivity. A reduction in the quality and quantity of resources available to herbivores due to desertification under changing climate can have knock-on consequences for predators and may ultimately disrupt trophic cascades (limited evidence, low agreement) (Rey et al. 2017651; Walther 2010652). Reduced resource availability may also compromise immune response to novel pathogens, with increased pathogen dispersal associated with dust storms (Zinabu et al. 2018653). Responses to desertification are species-specific and mechanistic models are not yet able to accurately predict individual species’ responses to the many factors associated with desertification (Fuller et al. 2016654).

3.4.2

Impacts on socio-economic systems

Combined impacts of desertification and climate change on socio-economic development in drylands are complex. Figure 3.9 schematically represents our qualitative assessment of the magnitudes and the uncertainties associated with these impacts on attainment of the SDGs in dryland areas (UN 2015655). The impacts of desertification and climate change are difficult to isolate from the effects of other socio-economic, institutional and political factors (Pradhan et al. 2017656). However, there is high confidence that climate change will exacerbate the vulnerability of dryland populations to desertification, and that the combination of pressures coming from climate change and desertification will diminish opportunities for reducing poverty, enhancing food and nutritional security, empowering women, reducing disease burden, and improving access to water and sanitation. Desertification is embedded in SDG 15 (Target 15.3) and climate change is under SDG 13. The high confidence and high magnitude impacts depicted for these SDGs (Figure 3.9) indicate that the interactions between desertification and climate change strongly affect the achievement of the targets of SDGs 13 and 15.3, pointing at the need for the coordination of policy actions on land degradation neutrality and mitigation and adaptation to climate change. The following subsections present the literature and assessments which serve as the basis for Figure 3.9.

Figure 3.9

Socio-economic impacts of desertification and climate change with the SDG framework.

Socio-economic impacts of desertification and climate change with the SDG framework.

3.4.2.1

Impacts on poverty

Climate change has a high potential to contribute to poverty particularly through the risks coming from extreme weather events (Olsson et al. 2014657). However, the evidence rigourously attributing changes in observed poverty to climate change impacts is currently not available. On the other hand, most of the research on links between poverty and desertification (or more broadly, land degradation) focused on whether or not poverty is a cause of land degradation (Gerber et al. 2014658; Vu et al. 2014659; Way 2016660) (Section 4.7.1). The literature measuring the extent to which desertification contributed to poverty globally is lacking: the related literature remains qualitative or correlational (Barbier and Hochard 2016661). At the local level, on the other hand, there is limited evidence and high agreement that desertification increased multidimensional poverty. For example, Diao and Sarpong (2011)662 estimated that land degradation lowered agricultural incomes in Ghana by 4.2 billion USD between 2006 and 2015, increasing the national poverty rate by 5.4% in 2015. Land degradation increased the probability of households becoming poor by 35% in Malawi and 48% in Tanzania (Kirui 2016663). Desertification in China was found to have resulted in substantial losses in income, food production and jobs (Jiang et al. 2014664). On the other hand, Ge et al. (2015)665 indicated that desertification was positively associated with growing incomes in Inner Mongolia in China in the short run since no costs were incurred for SLM, while in the long run higher incomes allowed allocation of more investments to reduce desertification. This relationship corresponds to the Environmental Kuznets Curve, which posits that environmental degradation initially rises and subsequently falls with rising income (e.g., Stern 2017666). There is limited evidence on the validity of this hypothesis regarding desertification.

3.4.2.2

Impacts on food and nutritional insecurity

About 821 million people globally were food insecure in 2017, of whom 63% in Asia, 31% in Africa and 5% in Latin America and the Caribbean (FAO et al. 2018667). The global number of food insecure people rose by 37 million since 2014. Changing climate variability, combined with a lack of climate resilience, was suggested as a key driver of this increase (FAO et al. 2018668). Sub-Saharan Africa, East Africa and South Asia had the highest share of undernourished populations in the world in 2017, with 28.8%, 31.4% and 33.7% respectively (FAO et al. 2018669). The major mechanism through which climate change and desertification affect food security is through their impacts on agricultural productivity. There is robust evidence pointing to negative impacts of climate change on crop yields in dryland areas (high agreement) (Hochman et al. 2017670; Nelson et al. 2010671; Zhao et al. 2017672) (Sections 3.4.1, 5.2.2 and 4.7.2). There is also robust evidence and high agreement on the losses in agricultural productivity and incomes due to desertification (Kirui 2016673; Moussa et al. 2016674; Mythili and Goedecke 2016675; Tun et al. 2015676). Nkonya et al. (2016a)677 estimated that cultivating wheat, maize, and rice with unsustainable land management practices is currently resulting in global losses of 56.6 billion USD annually, with another 8.7 billion USD of annual losses due to lower livestock productivity caused by rangeland degradation. However, the extent to which these losses affected food insecurity in dryland areas is not known. Lower crop yields and higher agricultural prices worsen existing food insecurity, especially for net food-buying rural households and urban dwellers. Climate change and desertification are not the sole drivers of food insecurity, but especially in the areas with high dependence on agriculture, they are among the main contributors.

3.4.2.3

Impacts on human health through dust storms

The frequency and intensity of dust storms are increasing due to land-use and land-cover changes and climate-related factors (Section 2.4) particularly in some regions of the world such as the Arabian Peninsula (Jish Prakash et al. 2015678; Yu et al. 2015679; Gherboudj et al. 2017680; Notaro et al. 2013681; Yu et al. 2013682; Alobaidi et al. 2017683; Maghrabi et al. 2011684; Almazroui et al. 2018685) and broader Middle East (Rashki et al. 2012686; Türkeş 2017687; Namdari et al. 2018688) as well as Central Asia (Indoitu et al. 2015689; Xi and Sokolik 2015690), with growing negative impacts on human health (high confidence) (Díaz et al. 2017691; Goudarzi et al. 2017692; Goudie 2014693; Samoli et al. 2011694). Dust storms transport particulate matter, pollutants, pathogens and potential allergens that are dangerous for human health over long distances (Goudie and Middleton 2006695; Sprigg 2016696). Particulate matter (PM; that is, the suspended particles in the air of up to 10 micrometres (PM10) or less in size), have damaging effects on human health (Díaz et al. 2017697; Goudarzi et al. 2017698; Goudie 2014699; Samoli et al. 2011700). The health effects of dust storms are largest in areas in the immediate vicinity of their origin, primarily the Sahara Desert, followed by Central and eastern Asia, the Middle East and Australia (Zhang et al. 2016701), however, there is robust evidence showing that the negative health effects of dust storms reach a much wider area (Bennett et al. 2006702; Díaz et al. 2017703; Kashima et al. 2016704; Lee et al. 2014705; Samoli et al. 2011706; Zhang et al. 2016707). The primary health effects of dust storms include damage to the respiratory and cardiovascular systems (Goudie 2013708). Dust particles with a diameter smaller than 2.5μm were associated with global cardiopulmonary mortality of about 402,000 people in 2005, with 3.47 million years of life lost in that single year (Giannadaki et al. 2014709). Although globally only 1.8% of cardiopulmonary deaths were caused by dust storms, in the countries of the Sahara region, Middle East, South and East Asia, dust storms were suggested to be the cause of 15–50% of all cardiopulmonary deaths (Giannadaki et al. 2014710). A 10 μgm-3 increase in PM10 dust particles was associated with mean increases in non-accidental mortality from 0.33% to 0.51% across different calendar seasons in China, Japan and South Korea (Kim et al. 2017711). The percentage of all-cause deaths attributed to fine particulate matter in Iranian cities affected by Middle Eastern dust storms (MED) was 0.56–5.02%, while the same percentage for non-affected cities was 0.16–4.13% (Hopke et al. 2018712). Epidemics of meningococcal meningitis occur in the Sahelian region during the dry seasons with dusty conditions (Agier et al. 2012713; Molesworth et al. 2003714). Despite a strong concentration of dust storms in the Sahel, North Africa, the Middle East and Central Asia, there is relatively little research on human health impacts of dust storms in these regions. More research on health impacts and related costs of dust storms, as well as on public health response measures, can help in mitigating these health impacts.

3.4.2.4

Impacts on gender equality

Environmental issues such as desertification and impacts of climate change have been increasingly investigated through a gender lens (Bose (n.d.)715; Broeckhoven and Cliquet 2015716; Kaijser and Kronsell 2014717; Kiptot et al. 2014718; Villamor and van Noordwijk 2016719). There is medium evidence and high agreement that women will be impacted more than men by environmental degradation (Arora-Jonsson 2011720; Gurung et al. 2006721) (Cross-Chapter Box 11 in Chapter 7). Socially structured gender-specific roles and responsibilities, daily activities, access and control over resources, decision-making and opportunities lead men and women to interact differently with natural resources and landscapes. For example, water scarcity affected women more than men in rural Ghana as they had to spend more time in fetching water, which has implications on time allocations for other activities (Ahmed et al. 2016722). Despite the evidence pointing to differentiated impact of environmental degradation on women and men, gender issues have been marginally addressed in many land restoration and rehabilitation efforts, which often remain gender-blind. Although there is robust evidence on the location-specific impacts of climate change and desertification on gender equality, there is limited evidence on the gender-related impacts of land restoration and rehabilitation activities. Women are usually excluded from local decision-making on actions regarding desertification and climate change. Socially constructed gender-specific roles and responsibilities are not static because they are shaped by other factors such as wealth, age, ethnicity and formal education (Kaijser and Kronsell 2014723; Villamor et al. 2014724). Hence, women’s and men’s environmental knowledge and priorities for restoration often differ (Sijapati Basnett et al. 2017725). In some areas where sustainable land options (e.g., agroforestry) are being promoted, women were not able to participate due to culturally embedded asymmetries in power relations between men and women (Catacutan and Villamor 2016726). Nonetheless women, particularly in the rural areas, remain heavily involved in securing food for their households. Food security for them is associated with land productivity and women’s contribution to address desertification is crucial.

3.4.2.5

Impacts on water scarcity and use

Reduced water retention capacity of degraded soils amplifies floods (de la Paix et al. 2011727), reinforces degradation processes through soil erosion, and reduces annual intake of water to aquifers, exacerbating existing water scarcities (Le Roux et al. 2017728; Cano et al. 2018729). Reduced vegetation cover and more intense dust storms were found to intensify droughts (Cook et al. 2009730). Moreover, secondary salinisation in the irrigated drylands often requires leaching with considerable amounts of water (Greene et al. 2016731; Wichelns and Qadir 2015732). Thus, different types of soil degradation increase water scarcity both through lower water quantity and quality (Liu et al. 2017733; Liu et al. 2016c734). All these processes reduce water availability for other needs. In this context, climate change will further intensify water scarcity in some dryland areas and increase the frequency of droughts (medium confidence) (IPCC 2013735; Zheng et al. 2018736) (Section 2.2). Higher water scarcity may imply growing use of wastewater effluents for irrigation (Pedrero et al. 2010737). The use of untreated wastewater exacerbates soil degradation processes (Tal 2016738; Singh et al. 2004739; Qishlaqi et al. 2008740; Hanjra et al. 2012741), in addition to negative human health impacts (Faour-Klingbeil and Todd 2018742; Hanjra et al. 2012743). Climate change will thus amplify the need for integrated land and water management for sustainable development.

3.4.2.6

Impacts on energy infrastructure through dust storms

Desertification leads to conditions that favour the production of dust storms (high confidence) (Section 3.3.1). There is robust evidence and high agreement that dust storms negatively affect the operational potential of solar and wind power harvesting equipment through dust deposition, reduced reach of solar radiation and increasing blade-surface roughness, and can also reduce effective electricity distribution in high-voltage transmission lines (Zidane et al. 2016744; Costa et al. 2016745; Lopez-Garcia et al. 2016746; Maliszewski et al. 2012747; Mani and Pillai 2010748; Mejia and Kleissl 2013749; Mejia et al. 2014750; Middleton 2017751; Sarver et al. 2013752; Kaufman et al. 2002753; Kok et al. 2018754). Direct exposure to desert dust storm can reduce energy generation efficiency of solar panels by 70–80% in one hour (Ghazi et al. 2014755). (Saidan et al. 2016756) indicated that in the conditions of Baghdad, Iraq, one month’s exposure to weather reduced the efficiency of solar modules by 18.74% due to dust deposition. In the Atacama desert, Chile, one month’s exposure reduced thin-film solar module performance by 3.7–4.8% (Fuentealba et al. 2015757). This has important implications for climate change mitigation efforts using the expansion of solar and wind energy generation in dryland areas for substituting fossil fuels. Abundant access to solar energy in many dryland areas makes them high-potential locations for the installation of solar energy generating infrastructure. Increasing desertification, resulting in higher frequency and intensity of dust storms imposes additional costs for climate change mitigation through deployment of solar and wind energy harvesting facilities in dryland areas. Most frequently used solutions to this problem involve physically wiping or washing the surface of solar devices with water. These result in additional costs and excessive use of already scarce water resources and labour (Middleton 2017758). The use of special coatings on the surface of solar panels can help prevent the deposition of dusts (Costa et al. 2016759; Costa et al. 2018760; Gholami et al. 2017761).

3.4.2.7

Impacts on transport infrastructure through dust storms and sand movement

Dust storms and movement of sand dunes often threaten the safety and operation of railway and road infrastructure in arid and hyper-arid areas, and can lead to road and airport closures due to reductions in visibility. For example, the dust storm on 10 March 2009 over Riyadh was assessed to be the strongest in the previous two decades in Saudi Arabia, causing limited visibility, airport shutdown and damages to infrastructure and environment across the city (Maghrabi et al. 2011762). There are numerous historical examples of how moving sand dunes led to the forced decommissioning of early railway lines built in Sudan, Algeria, Namibia and Saudi Arabia in the late 19th and early 20th century (Bruno et al. 2018763). Currently, the highest concentrations of railways vulnerable to sand movements are located in north-western China, Middle East and North Africa (Bruno et al. 2018764; Cheng and Xue 2014765). In China, sand dune movements are periodically disrupting the railway transport on the Linhai–Ceke line in north-western China and on the Lanzhou–Xinjiang High-speed Railway in western China, with considerable clean-up and maintenance costs (Bruno et al. 2018766; Zhang et al. 2010767). There are large-scale plans for expansion of railway networks in arid areas of China, Central Asia, North Africa, the Middle East, and eastern Africa. For example, The Belt and Road Initiative promoted by China, the Gulf Railway project by the Cooperation Council for the Arab States of the Gulf or Lamu Port–South Sudan–Ethiopia Transport (LAPSSET) Corridor in Eastern Africa. These investments have long-term return and operation periods. Their construction and associated engineering solutions will therefore benefit from careful consideration of potential desertification and climate change effects on sand storms and dune movements.

3.4.2.8

Impacts on conflicts

There is low confidence in climate change and desertification leading to violent conflicts. There is medium evidence and low agreement that climate change and desertification contribute to already existing conflict potentials (Herrero 2006768; von Uexkull et al. 2016769; Theisen 2017770; Olsson 2017771; Wischnath and Buhaug 2014772) (Section 4.7.3). To illustrate, Hsiang et al. (2013)773 found that each one standard deviation increase in temperature or rainfall was found to increase interpersonal violence by 4% and intergroup conflict by 14% (Hsiang et al. 2013774). However, this conclusion was disputed by Buhaug et al. (2014)775, who found no evidence linking climate variability to violent conflict after replicating Hsiang et al. (2013)776 by studying only violent conflicts. Almer et al. (2017)777 found that a one standard deviation increase in dryness raised the likelihood of riots in Sub-Saharan African countries by 8.3% during the 1990–2011 period. On the other hand, Owain and Maslin (2018)778 found that droughts and heatwaves were not significantly affecting the level of regional conflict in East Africa. Similarly, it was suggested that droughts and desertification in the Sahel played a relatively minor role in the conflicts in the Sahel in the 1980s, with the major reasons for the conflicts during this period being political, especially the marginalisation of pastoralists (Benjaminsen 2016779), corruption and rent-seeking (Benjaminsen et al. 2012780). Moreover, the role of environmental factors as the key drivers of conflicts was questioned in the case of Sudan (Verhoeven 2011781) and Syria (De Châtel 2014782). Selection bias, when the literature focuses on the same few regions where conflicts occurred and relates them to climate change, is a major shortcoming, as it ignores other cases where conflicts did not occur (Adams et al. 2018783) despite degradation of the natural resource base and extreme weather events.

3.4.2.9

Impacts on migration

Environmentally induced migration is complex and accounts for multiple drivers of mobility as well as other adaptation measures undertaken by populations exposed to environmental risk (high confidence). There is medium evidence and low agreement that climate change impacts migration. The World Bank (2018)784 predicted that 143 million people would be forced to move internally by 2050 if no climate action is taken. Focusing on asylum seekers alone, rather than the total number of migrants, Missirian and Schlenker (2017)785 predict that asylum applications to the European Union will increase from 28% (98,000 additional asylum applications per year) up to 188% (660,000 additional applications per year) depending on the climate scenario by 2100. While the modelling efforts have greatly improved over the years (Hunter et al. 2015786; McLeman 2011787; Sherbinin and Bai 2018788) and in particular, these recent estimates provide an important insight into potential future developments, the quantitative projections are still based on the number of people exposed to risk rather than the number of people who would actually engage in migration as a response to this risk (Gemenne 2011789; McLeman 2013790) and they do not take into account individual agency in migration decision nor adaptive capacities of individuals (Hartmann 2010791; Kniveton et al. 2011792; Piguet 2010793) (see Section 3.6.2 discussing migration as a response to desertification). Accordingly, the available micro-level evidence suggests that climate-related shocks are one of the many drivers of migration (Adger et al. 2014794; London Government Office for Science and Foresight 2011795; Melde et al. 2017796), but the individual responses to climate risk are more complex than commonly assumed (Gray and Mueller 2012a797). For example, despite strong focus on natural disasters, neither flooding (Gray and Mueller 2012b798; Mueller et al. 2014799) nor earthquakes (Halliday 2006800) were found to induce long-term migration; but instead, slow-onset changes, especially those provoking crop failures and heat stress, could affect household or individual migration decisions (Gray and Mueller 2012a801; Missirian and Schlenker 2017802; Mueller et al. 2014803). Out-migration from drought-prone areas has received particular attention (de Sherbinin et al. 2012804; Ezra and Kiros 2001805). A substantial body of literature suggests that households engage in local or internal migration as a response to drought (Findlay 2011806; Gray and Mueller 2012a807), while international migration decreases with drought in some contexts (Henry et al. 2004808), but might increase in contexts where migration networks are well established (Feng et al. 2010809; Nawrotzki and DeWaard 2016810; Nawrotzki et al. 2015811, 2016812). Similarly, the evidence is not conclusive with respect to the effect of environmental drivers, in particular desertification, on mobility. While it has not consistently entailed out-migration in the case of Ecuadorian Andes (Gray 2009, 2010813), environmental and land degradation increased mobility in Kenya and Nepal (Gray 2011814; Massey et al. 2010815), but marginally decreased mobility in Uganda (Gray 2011816). These results suggest that in some contexts, environmental shocks actually undermine households’ financial capacity to undertake migration (Nawrotzki and Bakhtsiyarava 2017817), especially in the case of the poorest households (Barbier and Hochard 2018818; Koubi et al. 2016819; Kubik and Maurel 2016820; McKenzie and Yang 2015821). Adding to the complexity, migration, especially to frontier areas, by increasing pressure on land and natural resources, might itself contribute to environmental degradation at the destination (Hugo 2008822; IPBES 2018a823; McLeman 2017824). The consequences of migration can also be salient in the case of migration to urban or peri-urban areas; indeed, environmentally induced migration can add to urbanisation (Section 3.6.2.2), often exacerbating problems related to poor infrastructure and unemployment.

3.4.2.10

Impacts on pastoral communities

Pastoral production systems occupy a significant portion of the world (Rass 2006825; Dong 2016826). Food insecurity among pastoral households is often high (Gomes 2006827) (Section 3.1.3). The Sahelian droughts of the 1970s–1980s provided an example of how droughts could affect livestock resources and crop productivity, contributing to hunger, out-migration and suffering for millions of pastoralists (Hein and De Ridder 2006828; Molua and Lambi 2007829). During these Sahelian droughts low and erratic rainfall exacerbated desertification processes, leading to ecological changes that forced people to use marginal lands and ecosystems. Similarly, the rate of rangeland degradation is now increasing because of environmental changes and overexploitation of resources (Kassahun et al. 2008830; Vetter 2005831). Desertification coupled with climate change is negatively affecting livestock feed and grazing species (Hopkins and Del Prado 2007832), changing the composition in favour of species with low forage quality, ultimately reducing livestock productivity (D’Odorico et al. 2013833; Dibari et al. 2016834) and increasing livestock disease prevalence (Thornton et al. 2009849). There is robust evidence and high agreement that weak adaptive capacity, coupled with negative effects from other climate-related factors, are predisposing pastoralists to increased poverty from desertification and climate change globally (López-i-Gelats et al. 2016835; Giannini et al. 2008836; IPCC 2007837). On the other hand, misguided policies such as enforced sedentarisation, and in certain cases protected area delineation (fencing), which restrict livestock mobility have hampered optimal use of grazing land resources (Du 2012838). Such policies have led to degradation of resources and out-migration of people in search of better livelihoods (Gebeye 2016839; Liao et al. 2015840). Restrictions on the mobile lifestyle are reducing the resilient adaptive capacity of pastoralists to natural hazards including extreme and variable weather conditions, drought and climate change (Schilling et al. 2014841). Furthermore, the exacerbation of the desertification phenomenon due to agricultural intensification (D’Odorico et al. 2013842) and land fragmentation caused by encroachment of agriculture into rangelands (Otuoma et al. 2009843; Behnke and Kerven 2013844) is threatening pastoral livelihoods. For example, commercial cotton (Gossypium hirsutum) production is crowding out pastoral systems in Benin (Tamou et al. 2018845). Food shortages and the urgency to produce enough crop for public consumption are leading to the encroachment of agriculture into productive rangelands and those converted rangelands are frequently prime lands used by pastoralists to produce feed and graze their livestock during dry years (Dodd 1994846). The sustainability of pastoral systems is therefore coming into question because of social and political marginalisation of those systems (Davies et al. 2016847) and also because of the fierce competition they are facing from other livelihood sources such as crop farming (Haan et al. 2016848).

3.5

Future projections

3.5.1

Future projections of desertification

Assessing the impact of climate change on future desertification is difficult as several environmental and anthropogenic variables interact to determine its dynamics. The majority of modelling studies regarding the future evolution of desertification rely on the analysis of specific climate change scenarios and Global Climate Models (GCMs) and their effect on a few processes or drivers that trigger desertification (Cross-Chapter Box 1 in Chapter 1).

With regards to climate impacts, the analysis of global and regional climate models concludes that under all representative concentration pathways (RCPs) potential evapotranspiration (PET) would increase worldwide as a consequence of increasing surface temperatures and surface water vapour deficit (Sherwood and Fu 2014850). Consequently, there would be associated changes in aridity indices that depend on this variable (high agreement, robust evidence) (Cook et al. 2014a851; Dai 2011852; Dominguez et al. 2010853; Feng and Fu 2013854; Ficklin et al. 2016855; Fu et al. 2016856; Greve and Seneviratne 1999857; Koutroulis 2019858; Scheff and Frierson 2015859). Due to the large increase in PET and decrease in precipitation over some subtropical land areas, aridity index will decrease in some drylands (Zhao and Dai 2015860), with one model estimating approximately 10% increase in hyper-arid areas globally (Zeng and Yoon 2009861). Increases in PET are projected to continue due to climate change (Cook et al. 2014a862; Fu et al. 2016863; Lin et al. 2015864; Scheff and Frierson 2015865). However, as noted in Sections 3.1.1 and 3.2.1, these PET calculations use assumptions that are not valid in an environment with changing CO2. Evidence from precipitation, runoff or photosynthetic uptake of CO2 suggest that a future warmer world will be less arid (Roderick et al. 2015866). Observations in recent decades indicate that the Hadley cell has expanded poleward in both hemispheres (Fu et al. 2006867; Hu and Fu 2007868; Johanson et al. 2009869; Seidel and Randel 2007870), and under all RCPs would continue expanding (Johanson et al. 2009871; Lu et al. 2007872). This expansion leads to the poleward extension of subtropical dry zones and hence an expansion in drylands on the poleward edge (Scheff and Frierson 2012873). Overall, this suggests that while aridity will increase in some places (high confidence), there is insufficient evidence to suggest a global change in dryland aridity (medium confidence).

Regional modelling studies confirm the outcomes of Global Climate Models (Africa: Terink et al. 2013874; China: Yin et al. 2015875; Brazil: Marengo and Bernasconi 2015876; Cook et al. 2012877; Greece: Nastos et al. 2013878; Italy: Coppola and Giorgi 2009879). According to the IPCC AR5 (IPCC 2013)880, decreases in soil moisture are detected in the Mediterranean, southwest USA and southern African regions. This is in line with alterations in the Hadley circulation and higher surface temperatures. This surface drying will continue to the end of this century under the RCP8.5 scenario (high confidence). Ramarao et al. (2015)881 showed that a future climate projection based on RCP4.5 scenario indicated the possibility for detecting the summer-time soil drying signal over the Indian region during the 21st century in response to climate change. The IPCC Special Report on Global Warming of 1.5°C (SR15) (Chapter 3; Hoegh-Guldberg et al. 2018882) concluded with ‘medium confidence’ that global warming by more than 1.5°C increases considerably the risk of aridity for the Mediterranean area and southern Africa. Miao et al. (2015b)883 showed an acceleration of desertification trends under the RCP8.5 scenario in the middle and northern part of Central Asia and some parts of north-western China. It is also useful to consider the effects of the dynamic–thermodynamical feedback of the climate. Schewe and Levermann (2017)884 show increases of up to 300% in the Central Sahel rainfall by the end of the century due to an expansion of the West African monsoon. Warming could trigger an intensification of monsoonal precipitation due to increases in ocean moisture availability.

The impacts of climate change on dust storm activity are not yet comprehensively studied and represent an important knowledge gap. Currently, GCMs are unable to capture recent observed dust emission and transport (Evan 2018885; Evan et al. 2014886), limiting confidence in future projections. Literature suggests that climate change decreases wind erosion/dust emission overall, with regional variation (low confidence). Mahowald et al. (2006)887 and Mahowald (2007)888 found that climate change led to a decrease in desert dust source areas globally using CMIP3 GCMs. Wang et al. (2009)889 found a decrease in sand dune movement by 2039 (increasing thereafter) when assessing future wind-erosion-driven desertification in arid and semi-arid China using a range of SRES scenarios and HadCM3 simulations. Dust activity in the Southern Great Plains in the USA was projected to increase, while in the Northern Great Plains it was projected to decrease under RCP8.5 climate change scenario (Pu and Ginoux 2017890). Evan et al. (2016)891 project a decrease in African dust emission associated with a slowdown of the tropical circulation in the high CO2 RCP8.5 scenario.

Global estimates of the impact of climate change on soil salinisation show that under the IS92a emissions scenario (a scenario prepared in 1992 that contains ‘business as usual’ assumptions) (Leggett et al. 1992892) the area at risk of salinisation would increase in the future (limited evidence, high agreement) (Schofield and Kirkby 2003893). Climate change has an influence on soil salinisation that induces further land degradation through several mechanisms that vary in their level of complexity. However, only a few examples can be found to illustrate this range of impacts, including the effect of groundwater table depletion (Rengasamy 2006894) and irrigation management (Sivakumar 2007895), salt migration in coastal aquifers with decreasing water tables (Sherif and Singh 1999896) (Section 4.10.7), and surface hydrology and vegetation that affect wetlands and favour salinisation (Nielsen and Brock 2009897).

3.5.1.1

Future vulnerability and risk of desertification

Following the conceptual framework developed in the Special Report on extreme events (SREX) (IPCC 2012898), future risks are assessed by examining changes in exposure (that is, presence of people; livelihoods; species or ecosystems; environmental functions, service, and resources; infrastructure; or economic, social or cultural assets; see Glossary), changes in vulnerability (that is, propensity or predisposition to be adversely affected; see Glossary) and changes in the nature and magnitude of hazards (that is, potential occurrence of a natural or human-induced physical event that causes damage; see Glossary). Climate change is expected to further exacerbate the vulnerability of dryland ecosystems to desertification by increasing PET globally (Sherwood and Fu 2014899). Temperature increases between 2°C and 4°C are projected in drylands by the end of the 21st century under RCP4.5 and RCP8.5 scenarios, respectively (IPCC 2013900). An assessment by Carrão et al. 2017901 showed an increase in drought hazards by late-century (2071–2099) compared to a baseline (1971–2000) under high RCPs in drylands around the Mediterranean, south-eastern Africa, and southern Australia. In Latin America, Morales et al. (2011)902 indicated that areas affected by drought will increase significantly by 2100 under SRES scenarios A2 and B2. The countries expected to be affected include Guatemala, El Salvador, Honduras and Nicaragua. In CMIP5 scenarios, Mediterranean types of climate are projected to become drier (Alessandri et al. 2014903; Polade et al. 2017904), with the equatorward margins being potentially replaced by arid climate types (Alessandri et al. 2014905). Globally, climate change is predicted to intensify the occurrence and severity of droughts (medium confidence) (Dai 2013906; Sheffield and Wood 2008907; Swann et al. 2016908; Wang 2005909; Zhao and Dai 2015910; Carrão et al. 2017911; Naumann et al. 2018912) (Section 2.2). Ukkola et al. (2018)913 showed large discrepancies between CMIP5 models for all types of droughts, limiting the confidence that can be assigned to projections of drought.

Drylands are characterised by high climatic variability. Climate impacts on desertification are not only defined by projected trends in mean temperature and precipitation values but are also strongly dependent on changes in climate variability and extremes (Reyer et al. 2013914). The responses of ecosystems depend on diverse vegetation types. Drier ecosystems are more sensitive to changes in precipitation and temperature (Li et al. 2018915; Seddon et al. 2016916; You et al. 2018917), increasing vulnerability to desertification. It has also been reported that areas with high variability in precipitation tend to have lower livestock densities and that those societies that have a strong dependence on livestock that graze natural forage are especially affected (Sloat et al. 2018918). Social vulnerability in drylands increases as a consequence of climate change that threatens the viability of pastoral food systems (Dougill et al. 2010919; López-i-Gelats et al. 2016920). Social drivers can also play an important role with regards to future vulnerability (Máñez Costa et al. 2011921). In the arid region of north-western China, Liu et al. (2016b)922 estimated that under RCP4.5 areas of increased vulnerability to climate change and desertification will surpass those with decreased vulnerability.

Using an ensemble of global climate, integrated assessment and impact models, Byers et al. (2018)923 investigated 14 impact indicators at different levels of global mean temperature change and socio-economic development. The indicators cover water, energy and land sectors. Of particular relevance to desertification are the water (e.g., water stress, drought intensity) and the land (e.g., habitat degradation) indicators. Under shared socio-economic pathway SSP2 (‘Middle of the Road’) at 1.5°C, 2°C and 3°C of global warming, the numbers of dryland populations exposed (vulnerable) to various impacts related to water, energy and land sectors (e.g., water stress, drought intensity, habitat degradation) are projected to reach 951 (178) million, 1152 (220) million and 1285 (277) million, respectively. While at global warming of 2°C, under SSP1 (‘Sustainability’), the exposed (vulnerable) dryland population is 974 (35) million, and under SSP3 (‘Fragmented World’) it is 1267 (522) million. Steady increases in the exposed and vulnerable populations are seen for increasing global mean temperatures. However much larger differences are seen in the vulnerable population under different SSPs. Around half the vulnerable population is in South Asia, followed by Central Asia, West Africa and East Asia.

3.5.2

Future projections of impacts

Future climate change is expected to increase the potential for increased soil erosion by water in dryland areas (medium confidence). Yang et al. (2003)924 use a Revised Universal Soil Loss Equation (RUSLE) model to study global soil erosion under historical, present and future conditions of both cropland and climate. Soil erosion potential has increased by about 17%, and climate change will increase this further in the future. In northern Iran, under the SRES A2 emission scenario the mean erosion potential is projected to grow by 45%, comparing the period 1991–2010 with 2031–2050 (Zare et al. 2016925).

A strong decrease in precipitation for almost all parts of Turkey was projected for the period 2021–2050 compared to 1971–2000 using Regional Climate Model, RegCM4.4 of the International Centre for Theoretical Physics (ICTP) under RCP4.5 and RCP8.5 scenarios (Türkeş et al. 2019926). The projected changes in precipitation distribution can lead to more extreme precipitation events and prolonged droughts, increasing Turkey’s vulnerability to soil erosion. In Portugal, a study comparing wet and dry catchments under A1B and B1 emission scenarios showed an increase in erosion in dry catchments (Serpa et al. 2015927). In Morocco an increase in sediment load is projected as a consequence of reduced precipitation (Simonneaux et al. 2015928). WGII AR5 concluded the impact of increases in heavy rainfall and temperature on soil erosion will be modulated by soil management practices, rainfall seasonality and land cover (Jiménez Cisneros et al. 2014929). Ravi et al. (2010)930 predicted an increase in hydrologic and aeolian soil erosion processes as a consequence of droughts in drylands. However, there are some studies that indicate that soil erosion will be reduced in Spain (Zabaleta et al. 2013931), Greece (Nerantzaki et al. 2015932) and Australia (Klik and Eitzinger 2010933), while others project changes in erosion as a consequence of the expansion of croplands (Borrelli et al. 2017934).

Potential dryland expansion implies lower carbon sequestration and higher risk of desertification (Huang et al. 2017935), with severe impacts on land usability and threats to food security. At the level of biomes (global-scale zones, generally defined by the type of plant life that they support in response to average rainfall and temperature patterns; see Glossary), soil carbon uptake is determined mostly by weather variability. The area of the land in which dryness controls CO2 exchange has risen by 6% since 1948 and is projected to expand by at least another 8% by 2050. In these regions net carbon uptake is about 27% lower than elsewhere (Yi et al. 2014936). Potential losses of soil carbon are projected to range from 9% to 12% of the total carbon stock in the 0–20 cm layer of soils in southern European Russia by end of this century (Ivanov et al. 2018937).

Desertification under climate change will threaten biodiversity in drylands (medium confidence). Rodriguez-Caballero et al. (2018)938 analysed the cover of biological soil crusts under current and future environmental conditions utilising an environmental niche modelling approach. Their results suggest that biological soil crusts currently cover approximately 1600 Mha in drylands. Under RCP scenarios 2.6 to 8.5, 25–40% of this cover will be lost by 2070 with climate and land use contributing equally. The predicted loss is expected to substantially reduce the contribution of biological soil crusts to nitrogen cycling (6.7–9.9 TgN yr−1) and carbon cycling (0.16–0.24 PgC yr−1) (Rodriguez-Caballero et al. 2018939). A study in Colorado Plateau, USA showed that changes in climate in drylands may damage the biocrust communities by promoting rapid mortality of foundational species (Rutherford et al. 2017940), while in the Southern California deserts climate change-driven extreme heat and drought may surpass the survival thresholds of some desert species (Bachelet et al. 2016941). In semi-arid Mediterranean shrublands in eastern Spain, plant species richness and plant cover could be reduced by climate change and soil erosion (García-Fayos and Bochet 2009942). The main drivers of species extinctions are land-use change, habitat pollution, over-exploitation, and species invasion, while climate change is indirectly linked to species extinctions (Settele et al. 2014943). Malcolm et al. (2006)944 found that more than 2000 plant species located within dryland biodiversity hotspots could become extinct within 100 years, starting 2004 (within the Cape Floristic Region, Mediterranean Basin and southwest Australia). Furthermore, it is suggested that land use and climate change could cause the loss of 17% of species within shrublands and 8% within hot deserts by 2050 (low confidence) (van Vuuren et al. 2006945). A study in the semi-arid Chinese Altai Mountains showed that mammal species richness will decline, rates of species turnover will increase, and more than 50% of their current ranges will be lost (Ye et al. 2018946).

Changing climate and land use have resulted in higher aridity and more droughts in some drylands, with the rising role of precipitation, wind and evaporation on desertification (Fischlin et al. 2007947). In a 2°C world, annual water discharge is projected to decline, and heatwaves are projected to pose risk to food production by 2070 (Waha et al. 2017948). However, Betts et al. (2018)949 found a mixed response of water availability (runoff) in dryland catchments to global temperature increases from 1.5°C to 2°C. The forecasts for Sub-Saharan Africa suggest that higher temperatures, increase in the number of heatwaves, and increasing aridity, will affect the rainfed agricultural systems (Serdeczny et al. 2017950). A study by Wang et al. (2009)951 in arid and semi-arid China showed decreased livestock productivity and grain yields from 2040 to 2099, threatening food security. In Central Asia, projections indicate a decrease in crop yields, and negative impacts of prolonged heat waves on population health (Reyer et al. 2017952) (Section 3.7.2). World Bank (2009)953 projected that, without the carbon fertilisation effect, climate change will reduce the mean yields for 11 major global crops – millet, field pea, sugar beet, sweet potato, wheat, rice, maize, soybean, groundnut, sunflower and rapeseed – by 15% in Sub-Saharan Africa, 11% in Middle East and North Africa, 18% in South Asia, and 6% in Latin America and the Caribbean by 2046–2055, compared to 1996–2005. A separate meta-analysis suggested a similar reduction in yields in Africa and South Asia due to climate change by 2050 (Knox et al. 2012954). Schlenker and Lobell (2010)955 estimated that in sub-Saharan Africa, crop production may be reduced by 17–22% due to climate change by 2050. At the local level, climate change impacts on crop yields vary by location (Section 5.2.2). Negative impacts of climate change on agricultural productivity contribute to higher food prices. The imbalance between supply and demand for agricultural products is projected to increase agricultural prices in the range of 31% for rice, to 100% for maize by 2050 (Nelson et al. 2010956), and cereal prices in the range between a 32% increase and a 16% decrease by 2030 (Hertel et al. 2010957). In southern European Russia, it is projected that the yields of grain crops will decline by 5–10% by 2050 due to the higher intensity and coverage of droughts (Ivanov et al. 2018958).

Climate change can have strong impacts on poverty in drylands (medium confidence) (Hallegatte and Rozenberg 2017959; Hertel and Lobell 2014960). Globally, Hallegatte et al. (2015)961 project that without rapid and inclusive progress on eradicating multidimensional poverty, climate change could increase the number of the people living in poverty by between 35 million and 122 million people by 2030. Although these numbers are global and not specific to drylands, the highest impacts in terms of the share of the national populations being affected are projected to be in the drylands areas of the Sahel region, eastern Africa and South Asia (Stephane Hallegatte et al. 2015962). The impacts of climate change on poverty vary depending on whether the household is a net agricultural buyer or seller. Modelling results showed that poverty rates would increase by about one-third among the urban households and non-agricultural self-employed in Malawi, Uganda, Zambia and Bangladesh due to high agricultural prices and low agricultural productivity under climate change (Hertel et al. 2010)963. On the contrary, modelled poverty rates fell substantially among agricultural households in Chile, Indonesia, the Philippines and Thailand, because higher prices compensated for productivity losses (Hertel et al. 2010964).

3.6

Responses to desertification under climate change

Achieving sustainable development of dryland livelihoods requires avoiding dryland degradation through SLM and restoring and rehabilitating the degraded drylands due to their potential wealth of ecosystem benefits and importance to human livelihoods and economies (Thomas 2008965). A broad suite of on-the-ground response measures exists to address desertification (Scholes 2009966), be it in the form of improved fire and grazing management, the control of erosion; integrated crop, soil and water management, among others (Liniger and Critchley 2007967; Scholes 2009968). These actions are part of the broader context of dryland development and long-term SLM within coupled socio-economic systems (Reynolds et al. 2007969; Stringer et al. 2017970; Webb et al. 2017971). Many of these response options correspond to those grouped under ‘land transitions’ in the IPCC Special Report on Global Warming of 1.5°C (Coninck et al. 2018972) (Table 6.4). It is therefore recognised that such actions require financial, institutional and policy support for their wide-scale adoption and sustainability over time (Sections 3.6.3, 4.8.5 and 6.4.4).

3.6.1

SLM technologies and practices: On-the-ground actions

A broad range of activities and measures can help avoid, reduce and reverse degradation across the dryland areas of the world. Many of these actions also contribute to climate change adaptation and mitigation, with further sustainable development co-benefits for poverty eradication and food security (high confidence) (Section 6.3). As preventing desertification is strongly preferable and more cost-effective than allowing land to degrade and then attempting to restore it (IPBES 2018b973; Webb et al. 2013974), there is a growing emphasis on avoiding and reducing land degradation, following the Land Degradation Neutrality framework (Cowie et al. 2018975; Orr et al. 2017976) (Section 4.8.5).

An assessment is made of six activities and measures practicable across the biomes and anthromes of the dryland domain (Figure 3.10). This suite of actions is not exhaustive, but rather a set of activities that are particularly pertinent to global dryland ecosystems. They are not necessarily exclusive to drylands and are often implemented across a range of biomes and anthromes (Figure 3.10; for afforestation, see Section 3.7.2, Cross-Chapter Box 2 in Chapter 1, and Chapter 4 (Section 4.8.3)). The use of anthromes as a structuring element for response options is based on the essential role of interactions between social and ecological systems in driving desertification within coupled socio-ecological systems (Cherlet et al. 2018977). The concept of the anthromes is defined in the Glossary and explored further in Chapters 1, 4 and 6.

The assessment of each action is twofold: firstly, to assess the ability of each action to address desertification and enhance climate change resilience, and secondly, to assess the potential impact of future climate change on the effectiveness of each action.

Figure 3.10

The typical distribution of on-the-ground actions across global biomes and anthromes.

The typical distribution of on-the-ground actions across global biomes and anthromes.

3.6.1.1

Integrated crop–soil–water management

Forms of integrated cropland management have been practiced in drylands for thousands of years (Knörzer et al. 2009978). Actions include planting a diversity of species including drought-resilient ecologically appropriate plants, reducing tillage, applying organic compost and fertiliser, adopting different forms of irrigation and maintaining vegetation and mulch cover. In the contemporary era, several of these actions have been adopted in response to climate change.

In terms of climate change adaptation, the resilience of agriculture to the impacts of climate change is strongly influenced by the underlying health and stability of soils as well as improvements in crop varieties, irrigation efficiency and supplemental irrigation, for example, through rainwater harvesting (medium evidence, high agreement) (Altieri et al. 2015979; Amundson et al. 2015980; Derpsch et al. 2010981; Lal 1997982; de Vries et al. 2012983). Desertification often leads to a reduction in ground cover that in turn results in accelerated water and wind erosion and an associated loss of fertile topsoil that can greatly reduce the resilience of agriculture to climate change (medium evidence, high agreement) (Touré et al. 2019984; Amundson et al. 2015985; Borrelli et al. 2017986; Pierre et al. 2017987). Amadou et al. (2011)988 note that even a minimal cover of crop residues (100 kg ha–1) can substantially decrease wind erosion.

Compared to conventional (flood or furrow) irrigation, drip irrigation methods are more efficient in supplying water to the plant root zone, resulting in lower water requirements and enhanced water use efficiency (robust evidence, high agreement) (Ibragimov et al. 2007989; Narayanamoorthy 2010990; Niaz et al. 2009991). For example, in the rainfed area of Fetehjang, Pakistan, the adoption of drip methods reduced water usage by 67–68% during the production of tomato, cucumber and bell peppers, resulting in a 68–79% improvement in water use efficiency compared to previous furrow irrigation (Niaz et al. 2009992). In India, drip irrigation reduced the amount of water consumed in the production of sugarcane by 44%, grapes by 37%, bananas by 29% and cotton by 45%, while enhancing yields by up to 29% (Narayanamoorthy 2010993). Similarly, in Uzbekistan, drip irrigation increased the yield of cotton by 10–19% while reducing water requirements by 18–42% (Ibragimov et al. 2007994).

A prominent response that addresses soil loss, health and cover is altering cropping methods. The adoption of intercropping (inter – and intra-row planting of companion crops) and relay cropping (temporally differentiated planting of companion crops) maintains soil cover over a larger fraction of the year, leading to an increase in production, soil nitrogen, species diversity and a decrease in pest abundance (robust evidence, medium agreement) (Altieri and Koohafkan 2008995; Tanveer et al. 2017996; Wilhelm and Wortmann 2004997). For example, intercropping maize and sorghum with Desmodium (an insect repellent forage legume) and Brachiaria (an insect trapping grass), which is being promoted in drylands of East Africa, led to a two-to-three-fold increase in maize production and an 80% decrease in stem boring insects (Khan et al. 2014998). In addition to changes in cropping methods, forms of agroforestry and shelterbelts are often used to reduce erosion and improve soil conditions (Section 3.7.2). For example, the use of tree belts of mixed species in northern China led to a reduction of surface wind speed and an associated reduction in soil temperature of up to 40% and an increase in soil moisture of up to 30% (Wang et al. 2008999).

A further measure that can be of increasing importance under climate change is rainwater harvesting (RWH), including traditional zai (small basins used to capture surface runoff), earthen bunds and ridges (Nyamadzawo et al. 20131001), fanya juus infiltration pits (Nyagumbo et al. 20191002), contour stone bunds (Garrity et al. 20101003) and semi-permeable stone bunds (often referred to by the French term digue filtrante) (Taye et al. 20151004). RWH increases the amount of water available for agriculture and livelihoods through the capture and storage of runoff, while at the same time reducing the intensity of peak flows following high-intensity rainfall events. It is therefore often highlighted as a practical response to dryness (i.e., long-term aridity and low seasonal precipitation) and rainfall variability, both of which are projected to become more acute over time in some dryland areas (Dile et al. 20131005; Vohland and Barry 20091006). For example, for drainage in Wadi Al-Lith, Saudi Arabia, the use of rainwater harvesting was suggested as a key climate change adaptation action (Almazroui et al. 20171007). There is robust evidence and high agreement that the implementation of RWH systems leads to an increase in agricultural production in drylands (Biazin et al. 20121008; Bouma and Wösten 20161009; Dile et al. 20131010). A global meta-analysis of changes in crop production due to the adoption of RWH techniques noted an average increase in yields of 78%, ranging from –28% to 468% (Bouma and Wösten 20161011). Of particular relevance to climate change in drylands is that the relative impact of RWH on agricultural production generally increases with increasing dryness. Relative yield improvements due to the adoption of RWH were significantly higher in years with less than 330 mm rainfall, compared to years with more than 330 mm (Bouma and Wösten 20161012). Despite delivering a clear set of benefits, there are some issues that need to be considered. The impact of RWH may vary at different temporal and spatial scales (Vohland and Barry 20091013). At a plot scale, RWH structures may increase available water and enhance agricultural production, SOC and nutrient availability, yet at a catchment scale, they may reduce runoff to downstream uses (Meijer et al. 20131014; Singh et al. 20121015; Vohland and Barry 20091016; Yosef and Asmamaw 20151017). Inappropriate storage of water in warm climes can lead to an increase in water related diseases unless managed correctly, for example, schistosomiasis and malaria (Boelee et al. 20131018).

Integrated crop–soil–water management may also deliver climate change mitigation benefits through avoiding, reducing and reversing the loss of SOC (Table 6.5). Approximately 20–30 Pg of SOC have been released into the atmosphere through desertification processes, for example, deforestation, overgrazing and conventional tillage (Lal 20041019). Activities, such as those associated with conservation agriculture (minimising tillage, crop rotation, maintaining organic cover and planting a diversity of species), reduce erosion, improve water use efficiency and primary production, increase inflow of organic material and enhance SOC over time, contributing to climate change mitigation and adaptation (high confidence) (Plaza-Bonilla et al. 20151020; Lal 20151021; Srinivasa Rao et al. 20151022; Sombrero and de Benito 20101023). Conservation agriculture practices also lead to increases in SOC (medium confidence). However, sustained carbon sequestration is dependent on net primary productivity and on the availability of crop-residues that may be relatively limited and often consumed by livestock or used elsewhere in dryland contexts (Cheesman et al. 20161024; Plaza-Bonilla et al. 20151025). For this reason, expected rates of carbon sequestration following changes in agricultural practices in drylands are relatively low (0.04–0.4 tC ha–1) and it may take a protracted period of time, even several decades, for carbon stocks to recover if lost (medium confidence) (Farage et al. 20071026; Hoyle et al. 20131027; Lal 20041028). This long recovery period enforces the rationale for prioritising the avoidance and reduction of land degradation and loss of C, in addition to restoration activities.

3.6.1.2

Grazing and fire management in drylands

Rangeland management systems such as sustainable grazing approaches and re-vegetation increase rangeland productivity (high confidence) (Table 6.5). Open grassland, savannah and woodland are home to the majority of world’s livestock production (Safriel et al. 20051029). Within these drylands areas, prevailing grazing and fire regimes play an important role in shaping the relative abundance of trees versus grasses (Scholes and Archer 19971030; Staver et al. 20111031; Stevens et al. 20171032), as well as the health of the grass layer in terms of primary production, species richness and basal cover (the propotion of the plant that is in the soil) (Plaza-Bonilla et al. 20151033; Short et al. 20031034). This in turn influences levels of soil erosion, soil nutrients, secondary production and additional ecosystem services (Divinsky et al. 20171035; Pellegrini et al. 20171036). A further set of drivers, including soil type, annual rainfall and changes in atmospheric CO2 may also define observed rangeland structure and composition (Devine et al. 20171037; Donohue et al. 20131038), but the two principal factors that pastoralists can manage are grazing and fire, by altering their frequency, type and intensity.

The impact of grazing and fire regimes on biodiversity, soil nutrients, primary production and further ecosystem services is not constant and varies between locations (Divinsky et al. 20171039; Fleischner 19941040; van Oijen et al. 20181041). Trade-offs may therefore need to be considered to ensure that rangeland diversity and production are resilient to climate change (Plaza-Bonilla et al. 20151042; van Oijen et al. 20181043). In certain locations, even light to moderate grazing has led to a significant decrease in the occurrence of particular species, especially forbs (O’Connor et al. 20111044; Scott-shaw and Morris 20151045). In other locations, species richness is only significantly impacted by heavy grazing and is able to withstand light to moderate grazing (Divinsky et al. 20171046). A context specific evaluation of how grazing and fire impact particular species may therefore be required to ensure the persistence of target species over time (Marty 20051047). A similar trade-off may need to be considered between soil carbon sequestration and livestock production. As noted by Plaza-Bonilla et al. (2015)1048 increasing grazing pressure has been found to increase SOC stocks in some locations, and decrease them in others. Where it has led to a decrease in soil carbon stocks, for example in Mongolia (Han et al. 20081049) and Ethiopia (Bikila et al. 20161050), trade-offs between carbon sequestration and the value of livestock to local livelihoods need be considered.

Although certain herbaceous species may be unable to tolerate grazing pressure, a complete lack of grazing or fire may not be desired in terms of ecosystems health. It can lead to a decrease in basal cover and the accumulation of moribund, unpalatable biomass that inhibits primary production (Manson et al. 20071051; Scholes 20091052). The utilisation of the grass sward through light to moderate grazing stimulates the growth of biomass and basal cover, and allows water services to be sustained over time (Papanastasis et al. 20171053; Scholes 20091054). Even moderate to heavy grazing in periods of higher rainfall may be sustainable, but constant heavy grazing during dry periods, and especially droughts, can lead to a reduction in basal cover, SOC, biological soil crusts, ecosystem services and an accelerated erosion (high agreement, robust evidence) (Archer et al. 20171055; Conant and Paustian 20031056; D’Odorico et al. 20131057; Geist and Lambin 20041058; Havstad et al. 20061059; Huang et al. 20071060; Manzano and Návar 20001061; Pointing and Belnap 20121062; Weber et al. 20161063). For this reason, the inclusion of drought forecasts and contingency planning in grazing and fire management programmes is crucial to avoid desertification (Smith and Foran 19921064; Torell et al. 20101065). It is an important component of avoiding and reducing early degradation. Although grasslands systems may be relatively resilient and can often recover from a moderately degraded state (Khishigbayar et al. 20151066; Porensky et al. 20161067), if a tipping point has been exceeded, restoration to a historic state may not be economical or ecologically feasible (D’Odorico et al. 20131068).

Together with livestock management (Table 6.5), the use of fire is an integral part of rangeland management, which can be applied to remove moribund and unpalatable forage, exotic weeds and woody species (Archer et al. 20171069). Fire has less of an effect on SOC and soil nutrients in comparison to grazing (Abril et al. 20051070), yet elevated fire frequency has been observed to lead to a decrease in soil carbon and nitrogen (Abril et al. 20051071; Bikila et al. 20161072; Bird et al. 20001073; Pellegrini et al. 20171074). Although the impact of climate change on fire frequency and intensity may not be clear due to its differing impact on fuel accumulation, suitable weather conditions and sources of ignition (Abatzoglou et al. 20181075; Littell et al. 20181076; Moritz et al. 20121077), there is an increasing use of prescribed fire to address several global change phenomena, for example, the spread of invasive species and bush encroachment, as well as the threat of intense runaway fires (Fernandes et al. 20131078; McCaw 20131079; van Wilgen et al. 20101080). Cross-Chapter Box 3 in Chapter 2 provides a further review of the interaction between fire and climate change.

There is often much emphasis on reducing and reversing the degradation of rangelands due to the wealth of benefits they provide, especially in the context of assisting dryland communities to adapt to climate change (Webb et al. 20171081; Woollen et al. 20161082). The emerging concept of ecosystem-based adaptation has highlighted the broad range of important ecosystem services that healthy rangelands can provide in a resilient manner to local residents and downstream economies (Kloos and Renaud 20161083; Reid et al. 20181084). In terms of climate change mitigation, the contribution of rangelands, woodland and sub-humid dry forest (e.g., Miombo woodland in south-central Africa) is often undervalued due to relatively low carbon stocks per hectare. Yet due to their sheer extent, the amount of carbon sequestered in these ecosystems is substantial and can make a valuable contribution to climate change mitigation (Lal 20041085; Pelletier et al. 20181086).

3.6.1.3

Clearance of bush encroachment

The encroachment of open grassland and savannah ecosystems by woody species has occurred for at least the past 100 years (Archer et al. 20171087; O’Connor et al. 20141088; Schooley et al. 20181089). Dependent on the type and intensity of encroachment, it may lead to a net loss of ecosystem services and be viewed as a form of desertification (Dougill et al. 20161090; O’Connor et al. 20141091). However, there are circumstances where bush encroachment may lead to a net increase in ecosystem services, especially at intermediate levels of encroachment, where the ability of the landscape to produce fodder for livestock is retained, while the production of wood and associated products increases (Eldridge et al. 20111092; Eldridge and Soliveres 20141093). This may be particularly important in regions such as southern Africa and India where over 65% of rural households depend on fuelwood from surrounding landscapes as well as livestock production (Komala and Prasad 20161094; Makonese et al. 20171095; Shackleton and Shackleton 20041096).

This variable relationship between the level of encroachment, carbon stocks, biodiversity, provision of water and pastoral value (Eldridge and Soliveres 20141097) can present a conundrum to policymakers, especially when considering the goals of three Rio Conventions: UNFCCC, UNCCD and UNCBD. Clearing intense bush encroachment may improve species diversity, rangeland productivity, the provision of water and decrease desertification, thereby contributing to the goals of the UNCBD and UNCCD as well as the adaptation aims of the UNFCCC. However, it would lead to the release of biomass carbon stocks into the atmosphere and potentially conflict with the mitigation aims of the UNFCCC.

For example, Smit et al. (2015)1098 observed an average increase in above-ground woody carbon stocks of 44 tC ha–1 in savannahs in northern Namibia. However, since bush encroachment significantly inhibited livestock production, there are often substantial efforts to clear woody species (Stafford-Smith et al. 20171099). Namibia has a national programme, currently in its early stages, aimed at clearing woody species through mechanical measures (harvesting of trees) as well as the application of arboricides (Smit et al. 20151100). However, the long-term success of clearance and subsequent improved fire and grazing management remains to be evaluated, especially restoration back towards an ‘original open grassland state’. For example, in northern Namibia, the rapid reestablishment of woody seedlings has raised questions about whether full clearance and restoration is possible (Smit et al. 20151101). In arid landscapes, the potential impact of elevated atmospheric CO2 (Donohue et al. 20131102; Kgope et al. 20101103) and opportunity to implement high-intensity fires that remove woody species and maintain rangelands in an open state has been questioned (Bond and Midgley 20001104). If these drivers of woody plant encroachment cannot be addressed, a new form of ‘emerging ecosystem’ (Milton 20031105) may need to be explored that includes both improved livestock and fire management as well as the utilisation of biomass as a long-term commodity and source of revenue (Smit et al. 20151106). Initial studies in Namibia and South Africa (Stafford-Smith et al. 20171107) indicate that there may be good opportunity to produce sawn timber, fencing poles, fuelwood and commercial energy, but factors such as the cost of transport can substantially influence the financial feasibility of implementation.

The benefit of proactive management that prevents land from being degraded (altering grazing systems or treating bush encroachment at early stages before degradation has been initiated) is more cost-effective in the long term and adds more resistance to climate change than treating lands after degradation has occurred (Webb et al. 20131108; Weltz and Spaeth 20121109). The challenge is getting producers to alter their management paradigm from short-term objectives to long-term objectives.

3.6.1.4

Combating sand and dust storms through sand dune stabilisation

Dust and sand storms have a considerable impact on natural and human systems (Sections 3.4.1 and 3.4.2). Application of sand dune stabilisation techniques contributes to reducing sand and dust storms (high confidence). Using a number of methods, sand dune stabilisation aims to avoid and reduce the occurrence of dust and sand storms (Mainguet and Dumay 20111110). Mechanical techniques include building palisades to prevent the movement of sand and reduce sand deposits on infrastructure. Chemical methods include the use of calcium bentonite or using silica gel to fix mobile sand (Aboushook et al. 20121111; Rammal and Jubair 20151112). Biological methods include the use of mulch to stabilise surfaces (Sebaa et al. 20151113; Yu et al. 20041114) and establishing permanent plant cover using pasture species that improve grazing at the same time (Abdelkebir and Ferchichi 20151115; Zhang et al. 20151116) (Section 3.7.1.3). When the dune is stabilised, woody perennials are introduced that are selected according to climatic and ecological conditions (FAO 20111117). For example, such re-vegetation processes have been implemented on the shifting dunes of the Tengger Desert in northern China leading to the stabilisation of sand and the sequestration of up to 10 tC ha–1 over a period of 55 years (Yang et al. 20141118).

3.6.1.5

Use of halophytes for the re-vegetation of saline lands

Soil salinity and sodicity can severely limit the growth and productivity of crops (Jan et al. 20171119) and lead to a decrease in available arable land. Leaching and drainage provides a possible solution, but can be prohibitively expensive. An alternative, more economical option, is the growth of halophytes (plants that are adapted to grow under highly saline conditions) that allow saline land to be used in a productive manner (Qadir et al. 20001120). The biomass produced can be used as forage, food, feed, essential oils, biofuel, timber, or fuelwood (Chughtai et al. 20151121; Mahmood et al. 20161122; Sharma et al. 20161123). A further co-benefit is the opportunity to mitigate climate change through the enhancement of terrestrial carbon stocks as land is re-vegetated (Dagar et al. 20141124; Wicke et al. 20131125). The combined use of salt-tolerant crops, improved irrigation practices, chemical remediation measures and appropriate mulch and compost is effective in reducing the impact of secondary salinisation (medium confidence).

In Pakistan, where about 6.2 Mha of agricultural land is affected by salinity, pioneering work on utilising salt-tolerant plants for the re-vegetation of saline lands (biosaline agriculture) was done in the early 1970s (NIAB 19971796). A number of local and exotic varieties were initially screened for salt tolerance in lab – and greenhouse-based studies, and then distributed to similar saline areas (Ashraf et al. 20101126). These included tree species (Acacia ampliceps, Acacia nilotica, Eucalyptus camaldulensis, Prosopis juliflora, Azadirachta indica) (Awan and Mahmood 20171127), forage plants (Leptochloa fusca, Sporobolus arabicus, Brachiaria mutica, Echinochloa sp., Sesbania and Atriplex spp.) and crop species including varieties of barley (Hordeum vulgare), cotton, wheat (Triticum aestivum) and Brassica spp. (Mahmood et al. 20161128) as well as fruit crops in the form of date palm (Phoenix dactylifera) that has high salt tolerance with no visible adverse effects on seedlings (Yaish and Kumar 20151129; Al-Mulla et al. 20131130; Alrasbi et al. 20101131). Pomegranate (Punica granatum L.) is another fruit crop of moderate to high salt tolerance. Through regulating growth form and nutrient balancing, it can maintain water content, chlorophyll fluorescence and enzyme activity at normal levels (Ibrahim 20161132; Okhovatian-Ardakani et al. 20101133).

In India and elsewhere, tree species including Prosopis juliflora, Dalbergia sissoo, and Eucalyptus tereticornis have been used to re-vegetate saline land. Certain biofuel crops in the form of Ricinus communis (Abideen et al. 20141134), Euphorbia antisyphilitica (Dagar et al. 20141135), Karelinia caspia (Akinshina et al. 20161797) and Salicornia spp. (Sanandiya and Siddhanta 20141136) are grown in saline areas, and Panicum turgidum (Koyro et al. 20131137) and Leptochloa fusca (Akhter et al. 20031138) have been grown as fodder crop on degraded soils with brackish water. In China, intense efforts are being made on the use of halophytes (Sakai et al. 20121139; Wang et al. 20181140). These examples reveal that there is great scope for saline areas to be used in a productive manner through the utilisation of halophytes. The most productive species often have yields equivalent to conventional crops, at salinity levels matching even that of seawater.

3.6.2

Socio-economic responses

Socio-economic and policy responses are often crucial in enhancing the adoption of SLM practices (Cordingley et al. 20151143; Fleskens and Stringer 20141144; Nyanga et al. 20161145) and for assisting agricultural households to diversify their sources of income (Barrett et al. 20171146; Shiferaw and Djido 20161147). Technology and socio-economic responses are not independent, but continuously interact.

3.6.2.1

Socio-economic responses for combating desertification under climate change

Desertification limits the choice of potential climate change mitigation and adaptation response options by reducing climate change adaptive capacities. Furthermore, many additional factors, for example, a lack of access to markets or insecurity of land tenure, hinder the adoption of SLM. These factors are largely beyond the control of individuals or local communities and require broader policy interventions (Section 3.6.3). Nevertheless, local collective action and ILK are still crucial to the ability of households to respond to the combined challenge of climate change and desertification. Raising awareness, capacity building and development to promote collective action and indigenous and local knowledge contribute to avoiding, reducing and reversing desertification under changing climate.

The use of indigenous and local knowledge enhances the success of SLM and its ability to address desertification (Altieri and Nicholls 20171148; Engdawork and Bork 20161149). Using indigenous and local knowledge for combating desertification could contribute to climate change adaptation strategies (Belfer et al. 20171150; Codjoe et al. 20141151; Etchart 20171152; Speranza et al. 20101153; Makondo and Thomas 20181154; Maldonado et al. 20161155; Nyong et al. 20071156). There are abundant examples of how indigenous and local knowledge, which are an important part of broader agroecological knowledge (Altieri 20181157), have allowed livelihood systems in drylands to be maintained despite environmental constraints. An example is the numerous traditional water harvesting techniques that are used across the drylands to adapt to dry spells and climate change. These include creating planting pits (zai, ngoro) and micro-basins, contouring hill slopes and terracing (Biazin et al. 20121158) (Section 3.6.1). Traditional ndiva water harvesting systems in Tanzania enable the capture of runoff water from highland areas to downstream community-managed micro-dams for subsequent farm delivery through small-scale canal networks (Enfors and Gordon 20081159). A further example are pastoralist communities located in drylands who have developed numerous methods to sustainably manage rangelands. Pastoralist communities in Morocco developed the agdal system of seasonally alternating use of rangelands to limit overgrazing (Dominguez 20141160) as well as to manage forests in the Moroccan High Atlas Mountains (Auclair et al. 20111161). Across the Arabian Peninsula and North Africa, a rotational grazing system, hema, was historically practiced by the Bedouin communities (Hussein 20111162; Louhaichi and Tastad 20101163). The Beni-Amer herders in the Horn of Africa have developed complex livestock breeding and selection systems (Fre 20181164). Although well adapted to resource-sparse dryland environments, traditional practices are currently not able to cope with increased demand for food and environmental changes (Enfors and Gordon 20081165; Engdawork and Bork 20161166). Moreover, there is robust evidence documenting the marginalisation or loss of indigenous and local knowledge (Dominguez 20141167; Fernández-Giménez and Fillat Estaque 20121168; Hussein 20111169; Kodirekkala 20171170; Moreno-Calles et al. 20121171). Combined use of indigenous and local knowledge and new SLM technologies can contribute to raising resilience to the challenges of climate change and desertification (high confidence) (Engdawork and Bork 20161172; Guzman et al. 20181173).

Collective action has the potential to contribute to SLM and climate change adaptation (medium confidence) (Adger 20031174; Engdawork and Bork 20161175; Eriksen and Lind 20091176; Ostrom 20091177; Rodima-Taylor et al. 20121178). Collective action is a result of social capital. Social capital is divided into structural and cognitive forms: structural corresponding to strong networks (including outside one’s immediate community); and cognitive encompassing mutual trust and cooperation within communities (van Rijn et al. 20121179; Woolcock and Narayan 20001180). Social capital is more important for economic growth in settings with weak formal institutions, and less so in those with strong enforcement of formal institutions (Ahlerup et al. 20091181). There are cases throughout the drylands showing that community by-laws and collective action successfully limited land degradation and facilitated SLM (Ajayi et al. 20161182; Infante 20171183; Kassie et al. 20131184; Nyangena 20081185; Willy and Holm-Müller 20131186; Wossen et al. 20151187). However, there are also cases when they did not improve SLM where they were not strictly enforced (Teshome et al. 20161188). Collective action for implementing responses to dryland degradation is often hindered by local asymmetric power relations and ‘elite capture’ (Kihiu 20161189; Stringer et al. 20071190). This illustrates that different levels and types of social capital result in different levels of collective action. In a sample of East, West and southern African countries, structural social capital in the form of access to networks outside one’s own community was suggested to stimulate the adoption of agricultural innovations, whereas cognitive social capital, associated with inward-looking community norms of trust and cooperation, was found to have a negative relationship with the adoption of agricultural innovations (van Rijn et al. 20121191). The latter is indirectly corroborated by observations of the impact of community-based rangeland management organisations in Mongolia. Although levels of cognitive social capital did not differ between them, communities with strong links to outside networks were able to apply more innovative rangeland management practices in comparison to communities without such links (Ulambayar et al. 20171192).

Farmer-led innovations. Agricultural households are not just passive adopters of externally developed technologies, but are active experimenters and innovators (Reij and Waters-Bayer 20011193; Tambo and Wünscher 20151194; Waters-Bayer et al. 20091195). SLM technologies co-generated through direct participation of agricultural households have higher chances of being accepted by them (medium confidence) (Bonney et al. 20161196; Vente et al. 20161197). Usually farmer-driven innovations are more frugal and better adapted to their resource scarcities than externally introduced technologies (Gupta et al. 20161198). Farmer-to-farmer sharing of their own innovations and mutual learning positively contribute to higher technology adoption rates (Dey et al. 20171199). This innovative ability can be given a new dynamism by combining it with emerging external technologies. For example, emerging low-cost phone applications (‘apps’) that are linked to soil and water monitoring sensors can provide farmers with previously inaccessible information and guidance (Cornell et al. 20131200; Herrick et al. 20171201; McKinley et al. 20171202; Steger et al. 20171203).

Currently, the adoption of SLM practices remains insufficient to address desertification and contribute to climate change adaptation and mitigation more extensively. This is due to the constraints on the use of indigenous and local knowledge and collective action, as well as economic and institutional barriers for SLM adoption (Banadda 20101204; Cordingley et al. 20151205; Lokonon and Mbaye 20181206; Mulinge et al. 20161207; Wildemeersch et al. 20151208) (Section 3.1.4.2; 3.6.3). Sustainable development of drylands under these socio-economic and environmental (climate change, desertification) conditions will also depend on the ability of dryland agricultural households to diversify their livelihoods sources (Boserup 19651209; Safriel and Adeel 20081210).

3.6.2.2

Socio-economic responses for economic diversification

Livelihood diversification through non-farm employment increases the resilience of rural households against desertification and extreme weather events by diversifying their income and consumption (high confidence). Moreover, it can provide the funds to invest into SLM (Belay et al. 20171211; Bryan et al. 20091212; Dumenu and Obeng 20161213; Salik et al. 20171214; Shiferaw et al. 20091215). Access to non-agricultural employment is especially important for poorer pastoral households as their small herd sizes make them less resilient to drought (Fratkin 20131216; Lybbert et al. 20041217). However, access to alternative opportunities is limited in the rural areas of many developing countries, especially for women and marginalised groups who lack education and social networks (Reardon et al. 20081218).

Migration is frequently used as an adaptation strategy to environmental change (medium confidence). Migration is a form of livelihood diversification and a potential response option to desertification and increasing risk to agricultural livelihoods under climate change (Walther et al. 20021219). Migration can be short-term (e.g., seasonal) or long-term, internal within a country or international. There is medium evidence showing rural households responding to desertification and droughts through all forms of migration, for example: during the Dust Bowl in the USA in the 1930s (Hornbeck 20121220); during droughts in Burkina Faso in the 2000s (Barbier et al. 20091221); in Mexico in the 1990s (Nawrotzki et al. 20161222); and by the Aymara people of the semi-arid Tarapacá region in Chile between 1820 and 1970, responding to declines in rainfall and growing demands for labour outside the region (Lima et al. 20161223). There is robust evidence and high agreement showing that migration decisions are influenced by a complex set of different factors, with desertification and climate change playing relatively lesser roles (Liehr et al. 20161224) (Section 3.4.2). Barrios et al. (2006)1225 found that urbanisation in Sub-Saharan Africa was partially influenced by climatic factors during the 1950–2000 period, in parallel to liberalisation of internal restrictions on labour movements: each 1% reduction in rainfall was associated with a 0.45% increase in urbanisation. This migration favoured more industrially diverse urban areas in Sub-Saharan Africa (Henderson et al. 20171226), because they offer more diverse employment opportunities and higher wages. Similar trends were also observed in Iran in response to water scarcity (Madani et al. 20161227).

However, migration involves some initial investments. For this reason, reductions in agricultural incomes due to climate change or desertification have the potential to decrease out-migration among the poorest agricultural households, who become less able to afford migration (Cattaneo and Peri 20161228), thus increasing social inequalities. There is medium evidence and high agreement that households with migrant worker members are more resilient against extreme weather events and environmental degradation compared to non-migrant households, who are more dependent on agricultural income (Liehr et al. 20161229; Salik et al. 20171230; Sikder and Higgins 20171231). Remittances from migrant household members potentially contribute to SLM adoptions, however, substantial out-migration was also found to constrain the implementation of labour-intensive land management practices (Chen et al. 20141232; Liu et al. 2016a1233).

3.6.3

Policy responses

The adoption of SLM practices depends on the compatibility of the technology with prevailing socio-economic and biophysical conditions (Sanz et al. 20171798). Globally, it was shown that every USD invested into restoring degraded lands yields social returns, including both provisioning and non-provisioning ecosystem services, in the range of 3–6 USD over a 30-year period (Nkonya et al. 2016a1234). A similar range of returns from land restoration activities was found in Central Asia (Mirzabaev et al. 20161235), Ethiopia (Gebreselassie et al. 20161236), India (Mythili and Goedecke 20161237), Kenya (Mulinge et al. 20161238), Niger (Moussa et al. 20161239) and Senegal (Sow et al. 20161240) (medium confidence). Despite these relatively high returns, there is robust evidence that the adoption of SLM practices remains low (Cordingley et al. 20151241; Giger et al. 20151242; Lokonon and Mbaye 20181243). Part of the reason for these low adoption rates is that the major share of the returns from SLM are social benefits, namely in the form of non-provisioning ecosystem services (Nkonya et al. 2016a1244). The adoption of SLM technologies does not always provide implementers with immediate private benefits (Schmidt et al. 20171245). High initial investment costs, institutional and governance constraints and a lack of access to technologies and equipment may inhibit their adoption further (Giger et al. 20151246; Sanz et al. 20171247; Schmidt et al. 20171248). However, not all SLM practices have high upfront costs. Analysing the World Overview of Conservation Approaches and Technologies (WOCAT) database, a globally acknowledged reference database for SLM, Giger et al. (2015)1249 found that the upfront costs of SLM technologies ranged from about 20 USD to 5000 USD, with the median cost being around 500 USD. Many SLM technologies are profitable within 3 to 10 years (medium confidence) (Djanibekov and Khamzina 20161250; Giger et al. 20151251; Moussa et al. 20161252; Sow et al. 20161253). About 73% of 363 SLM technologies evaluated were reported to become profitable within three years, while 97% were profitable within 10 years (Giger et al. 20151254). Similarly, it was shown that social returns from investments in restoring degraded lands will exceed their costs within six years in many settings across drylands (Nkonya et al. 2016a1255). However, even with affordable upfront costs, market failures – in the form of lack of access to credit, input and output markets, and insecure land tenure (Section 3.1.3) – result in the lack of adoption of SLM technologies (Moussa et al. 20161256). Payments for ecosystem services, subsidies for SLM, and encouragement of community collective action can lead to a higher level of adoption of SLM and land restoration activities (medium confidence) (Bouma and Wösten 20161257; Lambin et al. 20141258; Reed et al. 20151259; Schiappacasse et al. 20121260; van Zanten et al. 20141261) (Section 3.6.3). Enabling the policy responses discussed in this section will contribute to overcoming these market failures.

Many socio-economic factors shaping individual responses to desertification typically operate at larger scales. Individual households and communities do not exercise control over these factors, such as land tenure insecurity, lack of property rights, lack of access to markets, availability of rural advisory services, and agricultural price distortions. These factors are shaped by national government policies and international markets. As is the case with socio-economic responses, policy responses are classified below in two ways: those which seek to combat desertification under changing climate; and those which seek to provide alternative livelihood sources through economic diversification. These options are mutually complementary and contribute to all the three hierarchical elements of the Land Degradation Neutrality (LDN) framework, namely, avoiding, reducing and reversing land degradation (Cowie et al. 20181262; Orr et al. 20171263) (Sections 4.8.5 and 7.4.5, and Table 7.2). An enabling policy environment is a critical element for the achievement of LDN (Chasek et al. 20191264). Implementation of LDN policies can contribute to climate change adaptation and mitigation (high confidence) (Sections 3.6.1 and 3.7.2).

3.6.3.1

Policy responses towards combating desertification under climate change

Policy responses to combat desertification take numerous forms (Marques et al. 20161265). Below we discuss major policy responses consistently highlighted in the literature in connection with SLM and climate change, because these response options were found to strengthen adaptation capacities and to contribute to climate change mitigation. They include improving market access, empowering women, expanding access to agricultural advisory services, strengthening land tenure security, payments for ecosystem services, decentralised natural resource management, investing into research and monitoring of desertification and dust storms, and investing into modern renewable energy sources.

Policies aiming at improving market access, that is the ability to access output and input markets at lower costs, help farmers and livestock producers earn more profit from their produce. Increased profits both motivate and enable them to invest more in SLM. Higher access to input, output and credit markets was consistently found as a major factor in the adoption of SLM practices in a wide number of settings across the drylands (medium confidence) (Aw-Hassan et al. 20161266; Gebreselassie et al. 20161267; Mythili and Goedecke 20161268; Nkonya and Anderson 20151269; Sow et al. 2016). Lack of access to credit limits adjustments and agricultural responses to the impacts of desertification under changing climate, with long-term consequences for the livelihoods and incomes, as was shown during the North American Dust Bowl of the 1930s (Hornbeck 20121271). Government policies aimed at improving market access usually involve constructing and upgrading rural–urban transportation infrastructure and agricultural value chains, such as investments into construction of local markets, abattoirs and cold storage warehouses, as well as post-harvest processing facilities (McPeak et al. 2006). However, besides infrastructural constraints, providing improved access often involves relieving institutional constraints to market access (Little 20101272), such as improved coordination of cross-border food safety and veterinary regulations (Ait Hou et al. 20151273; Keiichiro et al. 20151274; McPeak et al. 2006; Unnevehr 20151275), and availability and access to market information systems (Bobojonov et al. 20161276; Christy et al. 20141277; Nakasone et al. 20141278).

Women’s empowerment. A greater emphasis on understanding gender-specific differences over land use and land management practices as an entry point can make land restoration projects more successful (medium confidence) (Broeckhoven and Cliquet 20151279; Carr and Thompson 20141280; Catacutan and Villamor 20161281; Dah-gbeto and Villamor 20161282). In relation to representation and authority to make decisions in land management and governance, women’s participation remains lacking particularly in the dryland regions. Thus, ensuring women’s rights means accepting women as equal members of the community and citizens of the state (Nelson et al. 20151283). This includes equitable access of women to resources (including extension services), networks, and markets. In areas where socio-cultural norms and practices devalue women and undermine their participation, actions for empowering women will require changes in customary norms, recognition of women’s (land) rights in government policies, and programmes to assure that their interests are better represented (Section 1.4.2 and Cross-Chapter Box 11 in Chapter 7). In addition, several novel concepts are recently applied for an in-depth understanding of gender in relation to science–policy interface. Among these are the concepts of intersectionality, that is, how social dimensions of identity and gender are bound up in systems of power and social institutions (Thompson-Hall et al. 20161284), bounded rationality for gendered decision-making, related to incomplete information interacting with limits to human cognition leading to judgement errors or objectively poor decision making (Villamor and van Noordwijk 20161285), anticipatory learning for preparing for possible contingencies and consideration of long-term alternatives (Dah-gbeto and Villamor 20161286) and systematic leverage points for interventions that produce, mark, and entrench gender inequality within communities (Manlosa et al. 20181287), which all aim to improve gender equality within agroecological landscapes through a systems approach.

Education and expanding access to agricultural services. Providing access to information about SLM practices facilitates their adoption (medium confidence) (Kassie et al. 20151288; Nkonya et al. 20151289; Nyanga et al. 20161291). Moreover, improving the knowledge of climate change, capacity building and development in rural areas can help strengthen climate change adaptive capacities (Berman et al. 20121292; Chen et al. 20181293; Descheemaeker et al. 20181294; Popp et al. 20091296; Tambo 20161297; Yaro et al. 20151298). Agricultural initiatives to improve the adaptive capacities of vulnerable populations were more successful when they were conducted through reorganised social institutions and improved communication, for example, in Mozambique (Osbahr et al. 20081299). Improved communication and education could be facilitated by wider use of new information and communication technologies (ICTs) (Peters et al. 20151300). Investments into education were associated with higher adoption of soil conservation measures, for example, in Tanzania (Tenge et al. 20041301). Bryan et al. (2009) found that access to information was the prominent facilitator of climate change adaptation in Ethiopia. However, resource constraints of agricultural services, and disconnects between agricultural policy and climate policy can hinder the dissemination of climate-smart agricultural technologies (Morton 20171302). Lack of knowledge was also found to be a significant barrier to implementation of soil rehabilitation programmes in the Mediterranean region (Reichardt 20101303). Agricultural services will be able to facilitate SLM best when they also serve as platforms for sharing indigenous and local knowledge and farmer innovations (Mapfumo et al. 20161304). Participatory research initiatives conducted jointly with farmers have higher chances of resulting in technology adoption (Bonney et al. 20161305; Rusike et al. 20061306; Vente et al. 2016). Moreover, rural advisory services are often more successful in disseminating technological innovations when they adopt commodity/value chain approaches, remain open to engagement in input supply, make use of new opportunities presented by ICTs, facilitate mutual learning between multiple stakeholders (Morton 20171307), and organise science and SLM information in a location-specific manner for use in education and extension (Bestelmeyer et al. 20171308).

Strengthening land tenure security. Strengthening land tenure security is a major factor contributing to the adoption of soil conservation measures in croplands (high confidence) (Bambio and Bouayad Agha 20181309; Higgins et al. 20181310; Holden and Ghebru 20161311; Paltasingh 20181312; Rao et al. 2016; Robinson et al. 20181313), thus contributing to climate change adaptation and mitigation. Moreover, land tenure security can lead to more investment in trees (Deininger and Jin 20061314; Etongo et al. 20151315). Land tenure recognition policies were found to lead to higher agricultural productivity and incomes, although with inter-regional variations, requiring an improved understanding of overlapping formal and informal land tenure rights (Lawry et al. 20171316). For example, secure land tenure increased investments into SLM practices in Ghana, but without affecting farm productivity (Abdulai et al. 20111317). Secure land tenure, especially for communally managed lands, helps reduce arbitrary appropriations of land for large-scale commercial farms (Aha and Ayitey 2017; Baumgartner 20171318; Dell’Angelo et al. 20171319). In contrast, privatisation of rangeland tenures in Botswana and Kenya led to the loss of communal grazing lands and actually increased rangeland degradation (Basupi et al. 20171320; Kihiu 20161321) as pastoralists needed to graze livestock on now smaller communal pastures. Since food insecurity in drylands is strongly affected by climate risks, there is robust evidence and high agreement that resilience to climate risks is higher with flexible tenure for allowing mobility for pastoralist communities, and not fragmenting their areas of movement (Behnke 19941323; Holden and Ghebru 20161324; Liao et al. 20171325; Turner et al. 20161326; Wario et al. 20161327). More research is needed on the optimal tenure mix, including low-cost land certification, redistribution reforms, market-assisted reforms and gender-responsive reforms, as well as collective forms of land tenure such as communal land tenure and cooperative land tenure (see Section 7.6.5 for a broader discussion of land tenure security under climate change).

Payment for ecosystem services (PES) provides incentives for land restoration and SLM (medium confidence) (Lambin et al. 20141328; Li et al. 2018; Reed et al. 20151329; Schiappacasse et al. 20121330). Several studies illustrate that the social costs of desertification are larger than its private cost (Costanza et al. 20141331; Nkonya et al. 2016a1332). Therefore, although SLM can generate public goods in the form of provisioning ecosystem services, individual land custodians underinvest in SLM as they are unable to reap these benefits fully. Payment for ecosystem services provides a mechanism through which some of these benefits can be transferred to land users, thereby stimulating further investment in SLM. The effectiveness of PES schemes depends on land tenure security and appropriate design, taking into account specific local conditions (Börner et al. 20171333). However, PES has not worked well in countries with fragile institutions (Karsenty and Ongolo 20121334). Equity and justice in distributing the payments for ecosystem services were found to be key for the success of the PES programmes in Yunnan, China (He and Sikor 2015). Yet, when reviewing the performance of PES programmes in the tropics, Calvet-Mir et al. (2015), found that they are generally effective in terms of environmental outcomes, despite being sometimes unfair in terms of payment distribution. It is suggested that the implementation of PES will be improved through decentralised approaches giving local communities a larger role in the decision-making process (He and Lang 2015).

Empowering local communities for decentralised natural resource management. Local institutions often play a vital role in implementing SLM initiatives and climate change adaptation measures (high confidence) (Gibson et al. 20051335; Smucker et al. 20151336). Pastoralists involved in community-based natural resource management in Mongolia had greater capacity to adapt to extreme winter frosts, resulting in less damage to their livestock (Fernandez-Gimenez et al. 20151337). Decreasing the power and role of traditional community institutions, due to top-down public policies, resulted in lower success rates in community-based programmes focused on rangeland management in Dirre, Ethiopia (Abdu and Robinson 20171338). Decentralised governance was found to lead to improved management in forested landscapes (Dressler et al. 20101339; Ostrom and Nagendra 20061340). However, there are also cases when local elites were placed in control and this decentralised natural resource management negatively impacted the livelihoods of the poorer and marginalised community members due to reduced access to natural resources (Andersson and Ostrom 20081341; Cullman 20151343; Dressler et al. 20101344).

The success of decentralised natural resource management initiatives depends on increased participation and empowerment of a diverse set of community members, not only local leaders and elites, in the design and management of local resource management institutions (Kadirbeyoglu and Özertan 20151345; Umutoni et al. 20161346), while considering the interactions between actors and institutions at different levels of governance (Andersson and Ostrom 20081347; Carlisle and Gruby 20171349; McCord et al. 20171351). An example of such programmes where local communities played a major role in land restoration and rehabilitation activities is the cooperative project on The National Afforestation and Erosion Control Mobilization Action Plan in Turkey, initiated by the Turkish Ministry of Agriculture and Forestry (Çalişkan and Boydak 20171352), with the investment of 1.8 billion USD between 2008 and 2012. The project mobilised local communities in cooperation with public institutions, municipalities, and non-governmental organisations, to implement afforestation, rehabilitation and erosion control measures, resulting in the afforestation and reforestation of 1.5 Mha (Yurtoglu 20151353). Moreover, some 1.75 Mha of degraded forest and 37,880 ha of degraded rangelands were rehabilitated. Finally, the project provided employment opportunities for 300,000 rural residents for six months every year, combining land restoration and rehabilitation activities with measures to promote socio-economic development in rural areas (Çalişkan and Boydak 20171354).

Investing in research and development. Desertification has received substantial research attention over recent decades (Turner et al. 20071355). There is also a growing research interest on climate change adaptation and mitigation interventions that help address desertification (Grainger 20091356). Agricultural research on SLM practices has generated a significant number of new innovations and technologies that increase crop yields without degrading the land, while contributing to climate change adaptation and mitigation (Section 3.6.1). There is robust evidence that such technologies help improve the food security of smallholder dryland farming households (Harris and Orr 20141357) (Section 6.3.5). Strengthening research on desertification is of high importance not only to meet SDGs but also to manage ecosystems effectively, based on solid scientific knowledge. More investment in research institutes and training the younger generation of researchers is needed for addressing the combined challenges of desertification and climate change (Akhtar-Schuster et al. 20111358; Verstraete et al. 20111359). This includes improved knowledge management systems that allow stakeholders to work in a coordinated manner by enhancing timely, targeted and contextualised information sharing (Chasek et al. 20111360). Knowledge and flow of knowledge on desertification is currently highly fragmented, constraining the effectiveness of those engaged in assessing and monitoring the phenomenon at various levels (Reed et al. 20111361). Improved knowledge and data exchange and sharing increase the effectiveness of efforts to address desertification (high confidence).

Developing modern renewable energy sources. Transitioning to renewable energy resources contributes to reducing desertification by lowering reliance on traditional biomass in dryland regions (medium confidence). This can also have socioeconomic and health benefits, especially for women and children (high confidence). Populations in most developing countries continue to rely on traditional biomass, including fuelwood, crop straws and livestock manure, for a major share of their energy needs, with the highest dependence in Sub-Saharan Africa (Amugune et al. 20171363; IEA 2013). Use of biomass for energy, mostly fuelwood (especially as charcoal), was associated with deforestation in some dryland areas (Iiyama et al. 20141364; Mekuria et al. 20181365; Neufeldt et al. 20151366; Zulu 20101367), while in some other areas there was no link between fuelwood collection and deforestation (Simon and Peterson 20181368; Swemmer et al. 20181369; Twine and Holdo 20161370). Moreover, the use of traditional biomass as a source of energy was found to have negative health effects through indoor air pollution (de la Sota et al. 20181371; Lim and Seow 2012), while also being associated with lower female labour force participation (Burke and Dundas 20151372). Jiang et al. (2014) indicated that providing improved access to alternative energy sources such as solar energy and biogas could help reduce the use of fuelwood in south-western China, thus alleviating the spread of rocky desertification. The conversion of degraded lands into cultivation of biofuel crops will affect soil carbon dynamics (Albanito et al. 20161374; Nair et al. 20111375) (Cross-Chapter Box 7 in Chapter 6). The use of biogas slurry as soil amendment or fertiliser can increase soil carbon (Galvez et al. 2012; Negash et al. 20171376). Large-scale installation of wind and solar farms in the Sahara Desert was projected to create a positive climate feedback through increased surface friction and reduced albedo, doubling precipitation over the neighbouring Sahel region with resulting increases in vegetation (Li et al. 2018). Transition to renewable energy sources in high-income countries in dryland areas primarily contributes to reducing GHG emissions and mitigating climate change, with some other co-benefits such as diversification of energy sources (Bang 20101377), while the impacts on desertification are less evident. The use of renewable energy has been proposed as an important mitigation option in dryland areas as well (El-Fadel et al. 20031378). Transitions to renewable energy are being promoted by governments across drylands (Cancino-Solórzano et al. 20161379; Hong et al. 20131380; Sen and Ganguly 2017) including in fossil-fuel rich countries (Farnoosh et al. 20141381; Dehkordi et al. 2017; Stambouli et al. 20121382; Vidadili et al. 20171383), despite important social, political and technical barriers to expanding renewable energy production (Afsharzade et al. 2016; Baker et al. 20141384; Elum and Momodu 20171385; Karatayev et al. 20161386). Improving social awareness about the benefits of transitioning to renewable energy resources, and access to hydro-energy, solar and wind energy contributes to their improved adoption (Aliyu et al. 20171387; Katikiro 2016).

Developing and strengthening climate services relevant for desertification. Climate services provide climate, drought and desertification-related information in a way that assists decision-making by individuals and organisations. Monitoring desertification, and integrating biogeophysical (climate, soil, ecological factors, biodiversity) and socio-economic (use of natural resources by local population) issues provide a basis for better vulnerability prediction and assessment (OSS, 2012; Vogt et al. 20111388). Examples of relevant services include: drought monitoring and early warning systems, often implemented by national climate and meteorological services but also encompassing regional and global systems (Pozzi et al. 20131389); and the Sand and Dust Storm Warning Advisory and Assessment System (SDS-WAS), created by WMO in 2007, in partnership with the World Health Organization (WHO) and the United Nations Environment Program (UNEP). Currently, there is also a lack of ecological monitoring in arid and semi-arid regions to study surface winds, dust and sand storms, and their impacts on ecosystems and human health (Bergametti et al. 20181390; Marticorena et al. 20101391). Reliable and timely climate services, relevant to desertification, can aid the development of appropriate adaptation and mitigation options, reducing the impact of desertification under changing climate on human and natural systems (high confidence) (Beegum et al. 20161392; Beegum et al. 2018; Cornet 20121393; Haase et al. 20181395; Sergeant et al. 20121396).

3.6.3.2

Policy responses supporting economic diversification

Despite policy responses for combating desertification, other factors will put strong pressures on the land, including climate change and growing food demands, as well as the need to reduce poverty and strengthen food security (Cherlet et al. 20181397) (Sections 6.1.4 and 7.2.2). Sustainable development of drylands and their resilience to combined challenges of desertification and climate change will thus also depend on the ability of governments to promote policies for economic diversification within agriculture and in non-agricultural sectors in order make dryland areas less vulnerable to desertification and climate change.

Investing into irrigation. Investments into expanding irrigation in dryland areas can help increase the resilience of agricultural production to climate change, improve labour productivity and boost production and income revenue from agriculture and livestock sectors (Geerts and Raes 20091399; Olayide et al. 20161400; Oweis and Hachum 20061401). This is particularly true for Sub-Saharan Africa, where currently only 6% of the cultivated areas are irrigated (Nkonya et al. 2016b1402). While renewable groundwater resources could help increase the share of irrigated land to 20.5–48.6% of croplands in the region (Altchenko and Villholth 20151403). On the other hand, over-extraction of groundwaters, mainly for irrigating crops, is becoming an important environmental problem in many dryland areas (Cherlet et al. 20181404), requiring careful design and planning of irrigation expansion schemes and use of water-efficient irrigation methods (Bjornlund et al. 20171405; Woodhouse et al. 20171406). For example, in Saudi Arabia, improving the efficiency of water management, for example through the development of aquifers, water recycling and rainwater harvesting, is part of a suite of policy actions to combat desertification (Bazza, et al. 20181407; Kingdom of Saudi Arabia 20161408). The expansion of irrigation to riverine areas, crucial for dry season grazing of livestock, needs to consider the income from pastoral activities, which is not always lower than income from irrigated crop production (Behnke and Kerven 20131409). Irrigation development could be combined with the deployment of clean-energy technologies in economically viable ways (Chandel et al. 20151410). For example, solar-powered drip irrigation was found to increase household agricultural incomes in Benin (Burney et al. 20101411). The sustainability of irrigation schemes based on solar-powered extraction of groundwaters depends on measures to avoid over-abstraction of groundwater resources and associated negative environmental impacts (Closas and Rap 20171412).

Expanding agricultural commercialisation. Faster poverty rate reduction and economic growth enhancement is realised when countries transition into the production of non-staple, high-value commodities and manage to build a robust agro-industry sector (Barrett et al. 20171413). Ogutu and Qaim (2019) found that agricultural commercialisation increased incomes and decreased multidimensional poverty in Kenya. Similar findings were earlier reported by Muriithi and Matz (2015) for commercialisation of vegetables in Kenya. Commercialisation of rice production was found to have increased smallholder welfare in Nigeria (Awotide et al. 20161414). Agricultural commercialisation contributed to improved household food security in Malawi, Tanzania and Uganda (Carletto et al. 20171415). However, such a transition did not improve farmers’ livelihoods in all cases (Reardon et al. 2009). High-value cash crop/animal production can be bolstered by wide-scale use of technologies, for example, mechanisation, application of inorganic fertilisers, crop protection and animal health products. Market oriented crop/animal production facilitates social and economic progress, with labour increasingly shifting out of agriculture into non-agricultural sectors (Cour 2001). Modernised farming, improved access to inputs, credit and technologies enhances competitiveness in local and international markets (Reardon et al. 20091417).

Facilitating structural transformations in rural economies implies that the development of non-agricultural sectors encourages the movement of labour from land-based livelihoods, vulnerable to desertification and climate change, to non-agricultural activities (Haggblade et al. 20101420). The movement of labour from agriculture to non-agricultural sectors is determined by relative labour productivities in these sectors (Shiferaw and Djido 20161421). Given already high underemployment in the farm sector, increasing labour productivity in the non-farm sector was found as the main driver of labour movements from farm sector to non-farm sector (Shiferaw and Djido 20161422). More investments into education can facilitate this process (Headey et al. 20141423). However, in some contexts, such as pastoralist communities in Xinjiang, China, income diversification was not found to improve the welfare of pastoral households (Liao et al. 20151424). Economic transformations also occur through urbanisation, involving the shift of labour from rural areas into gainful employment in urban areas (Jedwab and Vollrath 20151425). The majority of world population will be living in urban centres in the 21st century and this will require innovative means of agricultural production with minimum ecological footprint and less dependence on fossil fuels (Revi and Rosenzweig 20131426), while addressing the demand of cities (see Section 4.9.1 for discussion on urban green infrastructure). Although there is some evidence of urbanisation leading to the loss of indigenous and local ecological knowledge, however, indigenous and local knowledge systems are constantly evolving, and are also being integrated into urban environments (Júnior et al. 20161427; Reyes-García et al. 20131429; van Andel and Carvalheiro 20131430). Urban areas are attracting an increasing number of rural residents across the developing world (Angel et al. 20111431; Cour 20011432; Dahiya 20121433). Urban development contributes to expedited agricultural commercialisation by providing market outlet for cash crops, high-value crops, and livestock products. At the same time, urbanisation also poses numerous challenges in the form of rapid urban sprawl and pressures on infrastructure and public services, unemployment and associated social risks, which have considerable implications on climate change adaptive capacities (Bulkeley 20131434; Garschagen and Romero-Lankao 20151435).

3.6.4

Limits to adaptation, maladaptation, and barriers for mitigation

Chapter 16 in the IPCC Fifth Assessment Report (AR5) (Klein et al. 20151799) discusses the existence of soft and hard limits to adaptation, highlighting that values and perspectives of involved agents are relevant to identify limits (Sections 4.8.5.1 and 7.4.9). In that sense, adaptation limits vary from place to place and are difficult to generalise (Barnett et al. 20151486; Dow et al. 20131800; Klein et al. 20151801). Currently, there is a lack of knowledge on adaptation limits and potential maladaptation to combined effects of climate change and desertification (see Section 4.8.6 for discussion on resilience, thresholds, and irreversible land degradation, also relevant for desertification). However, the potential for residual risks (those risks which remain after adaptation efforts were taken, irrespective of whether they are tolerable or not, tolerability being a subjective concept) and maladaptive outcomes is high (high confidence). Some examples of residual risks are illustrated below in this section. Although SLM measures can help lessen the effects of droughts, they cannot fully prevent water stress in crops and resulting lower yields (Eekhout and de Vente 20191487). Moreover, although in many cases SLM measures can help reduce and reverse desertification, there would still be short-term losses in land productivity. Irreversible forms of land degradation (for example, loss of topsoil, severe gully erosion) can lead to the complete loss of land productivity. Even when solutions are available, their costs could be prohibitive, presenting the limits to adaptation (Dixon et al. 20131488). If warming in dryland areas surpasses human thermal physiological thresholds (Klein et al. 2015; Waha et al. 20131489), adaptation could eventually fail (Kamali et al. 20181490). Catastrophic shifts in ecosystem functions and services (for example coastal erosion (Chen et al. 2015; Schneider and Kéfi 20161491) (Section 4.9.8)) and economic factors can also result in adaptation failure (Evans et al. 2015). Despite the availability of numerous options that contribute to combating desertification, climate change adaptation and mitigation, there are also chances of maladaptive actions (medium confidence) (see Glossary). Some activities favouring agricultural intensification in dryland areas can become maladaptive due to their negative impacts on the environment (medium confidence). Agricultural expansion to meet food demands can come through deforestation and consequent diminution of carbon sinks (Godfray and Garnett 20141492; Stringer et al. 20121493). Agricultural insurance programmes encouraging higher agricultural productivity and measures for agricultural intensification can result in detrimental environmental outcomes in some settings (Guodaar et al. 20191494; Müller et al. 20171495) (Table 6.12). Development of more drought-tolerant crop varieties is considered as a strategy for adaptation to shortening rainy seasons, but this can also lead to a loss of local varieties (Al Hamndou and Requier-Desjardins 20081496). Livelihood diversification to collecting and selling firewood and charcoal production can exacerbate deforestation (Antwi-Agyei et al. 20181497). Avoiding maladaptive outcomes can often contribute both to reducing the risks from climate change and combating desertification (Antwi-Agyei et al. 20181498). Avoiding, reducing and reversing desertification would enhance soil fertility, increase carbon storage in soils and biomass, thus reducing carbon emissions from soils to the atmosphere (Section 3.7.2 and Cross-Chapter Box 2 in Chapter 1). In specific locations, there may be barriers for some of these activities. For example, afforestation and reforestation programmes can contribute to reducing sand storms and increasing carbon sinks in dryland regions (Chu et al. 2019) (Sections 3.6.1 and 3.7.2). However, implementing agroforestry measures in arid locations can be constrained by lack of water (Apuri et al. 20181499), leading to a trade-off between soil carbon sequestration and other water uses (Cao et al. 2018). Thus, even when solutions are available, social, economic and institutional constraints could post barriers to their implementation (medium confidence).

3.7

Hotspots and case studies

The challenges of desertification and climate change in dryland areas across the world often have very location-specific characteristics. The five case studies in this section present rich experiences and lessons learnt on: (i) soil erosion, (ii) afforestation and reforestation through ‘green walls’, (iii) invasive plant species, (iv) oases in hyper-arid areas, and (v) integrated watershed management. Although it is impossible to cover all hotspots of desertification and on-the-ground actions from all dryland areas, these case studies present a more focused assessment of these five issues, which emerged as salient in the group discussions and several rounds of review of this chapter. The choice of these case studies was also motivated by the desire to capture a wide diversity of dryland settings.

3.7.1

Climate change and soil erosion

3.7.1.1

Soil erosion under changing climate in drylands

Soil erosion is a major form of desertification occurring in varying degrees in all dryland areas across the world (Section 3.2), with negative effects on dryland ecosystems (Section 3.4). Climate change is projected to increase soil erosion potential in some dryland areas through more frequent heavy rainfall events and rainfall variability (see Section 3.5.2 for a more detailed assessment) (Achite and Ouillon 20071500; Megnounif and Ghenim 20161501; Vachtman et al. 20131502; Zhang and Nearing 20051503). There are numerous soil conservation measures that can help reduce soil erosion (Section 3.6.1). Such soil management measures include afforestation and reforestation activities, rehabilitation of degraded forests, erosion control measures, prevention of overgrazing, diversification of crop rotations, and improvement in irrigation techniques, especially in sloping areas (Anache et al. 20181504; ÇEMGM 2017; Li and Fang 2016; Poesen 20181505; Ziadat and Taimeh 20131506). Effective measures for soil conservation can also use spatial patterns of plant cover to reduce sediment connectivity, and the relationships between hillslopes and sediment transfer in eroded channels (García-Ruiz et al. 20171507). The following three examples present lessons learnt from the soil erosion problems and measures to address them in different settings of Chile, Turkey and the Central Asian countries.

3.7.1.2

No-till practices for reducing soil erosion in central Chile

Soil erosion by water is an important problem in Chile. National assessments conducted in 1979, which examined 46% of the continental surface of the country, concluded that very high levels of soil erosion affected 36% of the territory. The degree of soil erosion increases from south to north. The leading locations in Chile are the region of Coquimbo with 84% of eroded soils (Lat. 29°S, semi-arid climate), the region of Valparaíso with 57% of eroded soils (Lat. 33°S, Mediterranean climate) and the region of O’Higgins with 37% of eroded soils (Lat. 34°S, Mediterranean climate). The most important drivers of soil erosion are soil, slope, climate erosivity (i.e., precipitation, intensity, duration and frequency) due to a highly concentrated rainy season, and vegetation structure and cover. In the region of Coquimbo, goat and sheep overgrazing have aggravated the situation (CIREN 20101508). Erosion rates reach up to 100 t ha–1 annually, having increased substantially over the last 50 years (Ellies 2000). About 10.4% of central Chile exhibits high erosion rates (greater than 1.1 t ha–1 annually) (Bonilla et al. 20101509).

Over the last few decades there has been an increasing interest in the development of no-till (also called zero tillage) technologies to minimise soil disturbance, reduce the combustion of fossil fuels and increase soil organic matter. No-till, in conjunction with the adoption of strategic cover crops, has positively impacted soil biology with increases in soil organic matter. Early evaluations by Crovetto, (1998) showed that no-till application (after seven years) had doubled the biological activity indicators compared to traditional farming and even surpassed those found in pasture (grown for the previous 15 years). Besides erosion control, additional benefits are an increase of water-holding capacity and reduction in bulk density. Currently, the above no-till farm experiment has lasted for 40 years and continues to report benefits to soil health and sustainable production (Reicosky and Crovetto 20141510). The influence of this iconic farm has resulted in the adoption of soil conservation practices – and especially no-till – in dryland areas of the Mediterranean climate region of central Chile (Martínez et al. 20111511). Currently, it has been estimated that the area under no-till farming in Chile varies between 0.13 and 0.2 Mha (Acevedo and Silva 20031512).

3.7.1.3

Combating wind erosion and deflation in Turkey: The greening desert of Karapınar

In Turkey, the amount of sediment recently released through erosion into seas was estimated to be 168 Mt yr-1 , which is considerably lower than the 500 Mt yr–1 that was estimated to be lost in the 1970s. The decrease in erosion rates is attributed to an increase in spatial extent of forests, rehabilitation of degraded forests, erosion control, prevention of overgrazing, and improvement in irrigation technologies. Soil conservation measures conducted in the Karapınar district, Turkey, exemplify these activities. The district is characterised by a semi-arid climate and annual average precipitation of 250–300 mm (Türkeş 20031513; Türkeş and Tatlı 20111514). In areas where vegetation was overgrazed or inappropriately tilled, the surface soil horizon was removed through erosion processes resulting in the creation of large drifting dunes that threatened settlements around Karapınar (Groneman 19681515). Such dune movement had begun to affect the Karapınar settlement in 1956 (Kantarcı et al. 20111516). Consequently, by the early 1960s, Karapınar town and nearby villages were confronted with the danger of abandonment due to out-migration in the early 1960s (Figure 3.11(1)). The reasons for increasing wind erosion in the Karapınar district can be summarised as follows: sandy material was mobilised following drying of the lake; hot and semi-arid climate conditions; overgrazing and use of pasture plants for fuel; excessive tillage; and strong prevailing winds.

Figure 3.11a

(1) A general view of a nearby village of Karapınar town in the early 1960s (Çarkaci 1999).

(1) A general view of a nearby village of Karapınar town in the early 1960s (Çarkaci 1999)1802.

Figure 3.11b

(2)A view of the Karapınar wind erosion area in 2013 (Photo: Murat Türkeş, 17 June 2019).

(2)A view of the Karapınar wind erosion area in 2013 (Photo: Murat Türkeş1803, 17 June 2019).

Figure 3.11c

(3) Construction of cane screens in the early 1960s in order to decrease wind speed and prevent movement of the sand accumulations and dunes; this was one of the physical measures during the prevention and mitigation period (Çarkaci 1999).

(3) Construction of cane screens in the early 1960s in order to decrease wind speed and prevent movement of the sand accumulations and dunes; this was one of the physical measures during the prevention and mitigation period (Çarkaci 19991804).

Figure 3.11d

(4) A view of mixed vegetation, which now covers most of the Karapınar wind erosion area in 2013, the main tree species of which were selected for afforestation with respect to their resistance to the arid continental climate conditions along with a warm/hot temperature regime over the district (Photo: Murat Türkeş, 17 June 2013).

(4) A view of mixed vegetation, which now covers most of the Karapınar wind erosion area in 2013, the main tree species of which were selected for afforestation with respect to their resistance to the arid continental climate conditions along with a warm/hot temperature regime over the district (Photo: Murat Türkeş1805, 17 June 2013).

Restoration and mitigation strategies were initiated in 1959, and today 4300 ha of land have been restored (Akay and Yildirim 20101517) (Figure 3.11 (2)), using specific measures: (i) physical measures: construction of cane screens to decrease wind speed and prevent sand movement (Figure 3.11(3)); (ii) restoration of cover: increasing grass cover between screens using seeds collected from local pastures or the cultivation of rye (Secale sp.) and wheat grass (Agropyron elongatum) that are known to grow in arid and hot conditions; and (iii) afforestation: saplings obtained from nursery gardens were planted and grown between these screens. Main tree species selected were oleaster (Eleagnus sp.), acacia (Robinia pseudeaccacia), ash (Fraxinus sp.), elm (Ulmus sp.) and maple (Acer sp.) (Figure 3.11 (4)). Economic growth occurred after controlling erosion and new tree nurseries have been established with modern irrigation. Potential negative consequences through the excessive use of water can be mitigated through engagement with local stakeholders and transdisciplinary learning processes, as well as by restoring the traditional land uses in the semi-arid Konya closed basin (Akça et al. 20161518).

3.7.1.4

Soil erosion in Central Asia under changing climate

Soil erosion is widely acknowledged to be a major form of degradation of Central Asian drylands, affecting a considerable share of croplands and rangelands. However, up-to-date information on the actual extent of eroded soils at the regional or country level is not available. The estimates compiled by Pender et al. (2009), based on the Central Asian Countries Initiative for Land Management (CACILM), indicate that about 0.8 Mha of the irrigated croplands were subject to high degree of soil erosion in Uzbekistan. In Turkmenistan, soil erosion was indicated to be occurring in about 0.7 Mha of irrigated land. In Kyrgyzstan, out of 1 Mha of irrigated land in the foothill zones, 0.76 Mha were subject to soil erosion by water, leading to losses in crop yields of 20–60% in these eroded soils. About 0.65 Mha of arable land were prone to soil erosion by wind (Mavlyanova et al. 20171519). Soil erosion is widespread in rainfed and irrigated areas in Kazakhstan (Saparov 2014). About 5 Mha of rainfed croplands were subject to high levels of soil erosion (Pender et al. 20091520). Soil erosion by water was indicated to be a major concern in sloping areas in Tajikistan (Pender et al. 20091521).

The major causes of soil erosion in Central Asia are related to human factors, primarily excessive water use in irrigated areas (Gupta et al. 20091522), deep ploughing and lack of maintenance of vegetative cover in rainfed areas (Suleimenov et al. 20141523), and overgrazing in rangelands (Mirzabaev et al. 20161524). Lack of good maintenance of watering infrastructure for migratory livestock grazing, and fragmentation of livestock herds led to overgrazing near villages, increasing the soil erosion by wind (Alimaev et al. 20081526). Overgrazing in the rangeland areas of the region (e.g., particularly in Kyzylkum) contributes to dust storms, coming primarily from the Ustyurt Plateau, desertified areas of Amudarya and Syrdarya rivers’ deltas, the dried seabed of the Aral Sea (now called Aralkum), and the Caspian Sea (Issanova and Abuduwaili 20171527; Xi and Sokolik 2015). Xi and Sokolik (2015) estimated that total dust emissions in Central Asia were 255.6 Mt in 2001, representing 10–17% of the global total.

Central Asia is one of the regions highly exposed to climate change, with warming levels projected to be higher than the global mean (Hoegh-Guldberg et al. 20181528), leading to more heat extremes (Reyer et al. 20171529). There is no clear trend in precipitation extremes, with some potential for moderate rise in occurrence of droughts. The diminution of glaciers is projected to continue in the Pamir and Tian Shan mountain ranges, a major source of surface waters along with seasonal snowmelt. Glacier melting will increase the hazards from moraine-dammed glacial lakes and spring floods (Reyer et al. 20171530). Increased intensity of spring floods creates favourable conditions for higher soil erosion by water, especially in the sloping areas in Kyrgyzstan and Tajikistan. The continuation of some of the current unsustainable cropland and rangeland management practices may lead to elevated rates of soil erosion, particularly in those parts of the region where climate change projections point to increases in floods (Kyrgyzstan, Tajikistan) or increases in droughts (Turkmenistan, Uzbekistan) (Hijioka et al. 20141531). Increasing water use to compensate for higher evapotranspiration due to rising temperatures and heat waves could increase soil erosion by water in the irrigated zones, especially in sloping areas and crop fields with uneven land levelling (Bekchanov et al. 20101532). The desiccation of the Aral Sea resulted in a hotter and drier regional microclimate, adding to the growing wind erosion in adjacent deltaic areas and deserts (Kust 19991533).

There are numerous sustainable land and water management practices available in the region for reducing soil erosion (Abdullaev et al. 20071534; Gupta et al. 20091535; Kust et al. 20141536; Nurbekov et al. 20161537). These include: improved land levelling and more efficient irrigation methods such as drip, sprinkler and alternate furrow irrigation (Gupta et al. 20091538); conservation agriculture practices, including no-till methods and maintenance of crop residues as mulch in the rainfed and irrigated areas (Kienzler et al. 20121539; Pulatov et al. 20121540); rotational grazing; institutional arrangements for pooling livestock for long-distance mobile grazing; reconstruction of watering infrastructure along the livestock migratory routes (Han et al. 2016; Mirzabaev et al. 20161541); afforesting degraded marginal lands (Djanibekov and Khamzina 20161543; Khamzina et al. 20091545; Khamzina et al. 20161546); integrated water resource management (Dukhovny et al. 20131547; Kazbekov et al. 20091548); and planting salt – and drought-tolerant halophytic plants as windbreaks in sandy rangelands (Akinshina et al. 20161549; Qadir et al. 20091550; Toderich et al. 20091551; Toderich et al. 20081552), and potentially the dried seabed of the former Aral Sea (Breckle 20131553). The adoption of enabling policies, such as those discussed in Section 3.6.3, can facilitate the adoption of these sustainable land and water management practices in Central Asia (high confidence) (Aw-Hassan et al. 20161554; Bekchanov et al. 20161555; Bobojonov et al. 20131556; Djanibekov et al. 20161557; Hamidov et al. 20161559; Mirzabaev et al. 20161560).

3.7.2

Green walls and green dams

This case study evaluates the experiences of measures and actions implemented to combat soil erosion, decrease dust storms, and to adapt to and mitigate climate change under the Green Wall and Green Dam programmes in East Asia (e.g., China) and Africa (e.g., Algeria, Sahara and the Sahel region). These measures have also been implemented in other countries, such as Mongolia (Do and Kang 2014; Lin et al. 2009), Turkey (Yurtoglu 20151562; Çalişkan and Boydak 20171563) and Iran (Amiraslani and Dragovich 20111564), and are increasingly considered as part of many national and international initiatives to combat desertification (Goffner et al. 20191565) (Cross-Chapter Box 2 in Chapter 1). Afforestation and reforestation programmes can contribute to reducing sand storms and increasing carbon sinks in dryland regions (high confidence). On the other hand, green wall and green dam programmes also decrease the albedo and hence increase the surface absorption of radiation, increasing the surface temperature. The net effect will largely depend on the balance between these and will vary from place to place depending on many factors.

3.7.2.1

The experiences of combating desertification in China

Arid and semi-arid areas of China, including north-eastern, northern and north-western regions, cover an area of more than 509 Mha, with annual rainfall of below 450 mm. Over the past several centuries, more than 60% of the areas in arid and semi-arid regions were used as pastoral and agricultural lands. The coupled impacts of past climate change and human activity have caused desertification and dust storms to become a serious problem in the region (Xu et al. 2010). In 1958, the Chinese government recognised that desertification and dust storms jeopardised the livelihoods of nearly 200 million people, and afforestation programmes for combating desertification have been initiated since 1978. China is committed to go beyond the Land Degradation Neutrality objective, as indicated by the following programmes that have been implemented. The Chinese Government began the Three North’s Forest Shelterbelt programme in Northeast China, North China, and Northwest China, with the goal to combat desertification and to control dust storms by improving forest cover in arid and semi-arid regions. The project is implemented in three stages (1978–2000, 2001–2020 and 2021–2050). In addition, the Chinese government launched the Beijing and Tianjin Sandstorm Source Treatment Project (2001–2010), Returning Farmlands to Forest Project (2003–present), and the Returning Grazing Land to Grassland Project (2003–present) to combat desertification, and for adaptation and mitigation of climate change (State Forestry Administration of China 20151566; Wang 20141567; Wang et al. 20131568).

The results of the fifth monitoring period (2010–2014) showed: (i) compared with 2009, the area of degraded land decreased by 12,120 km2 over a five-year period; (ii) in 2014, the average coverage of vegetation in the sand area was 18.33%, an increase of 0.7% compared with 17.63% in 2009, and the carbon sequestration increased by 8.5%; (iii) compared with 2009, the amount of wind erosion decreased by 33%, the average annual occurrence of sandstorms decreased by 20.3% in 2014; (iv) as of 2014, 203,700 km2 of degraded land were effectively managed, accounting for 38.4% of the 530,000 km2 of manageable desertified land; (v) the restoration of degraded land has created an annual output of 53.63 Mt of fresh and dried fruits, accounting for 33.9% of the total national annual output of fresh and dried fruits (State Forestry Administration of China 20151570). This has become an important pillar for economic development and a high priority for peasants as a method to eradicate poverty (State Forestry Administration of China 20151571).

Stable investment mechanisms for combating desertification have been established along with tax relief policies and financial support policies for guiding the country in its fight against desertification. The investments in scientific and technological innovation for combating desertification have been improved, the technologies for vegetation restoration under drought conditions have been developed, the popularisation and application of new technologies has been accelerated, and the training of technicians to assist farmers and herdsmen has been strengthened. To improve the monitoring capability and technical level of desertification studies, the monitoring network system has been strengthened, and the popularisation and application of modern technologies have been intensified (e.g., information technology and remote sensing) (Wu et al. 2015). Special laws on combating desertification have been decreed by the government. The provincial government’s responsibilities for desertification prevention and controlling objectives and laws have been strictly implemented.

Many studies showed that these projects generally played an active role in combating desertification and fighting against dust storms in China over the past several decades (high confidence) (Cao et al. 2018; State Forestry Administration of China 2015; Wang et al. 20131573; Wang et al. 20141574; Yang et al. 20131576). At the beginning of the projects, some problems appeared in some places due to lack of enough knowledge and experience (low confidence) (Jiang 20161578; Wang et al. 20101579). For example, some tree species selected were not well suited to local soil and climatic conditions (Zhu et al. 2007), and there was inadequate consideration of the limitation of the amount of available water on the carrying capacity of trees in some arid regions (Dai 2011; Feng et al. 20161580) (Section 3.6.4). In addition, at the beginning of the projects, there was an inadequate consideration of the effects of climate change on combating desertification (Feng et al. 20151581; Tan and Li 2015). Indeed, climate change and human activities over past years have influenced the desertification and dust storm control effects in China (Feng et al. 20151582; Wang et al. 20091583; Tan and Li 2015), and future climate change will bring new challenges for combating desertification in China (Wang et al. 20171584; Yin et al. 2015; Xu et al. 2019). In particular, the desertification risk in China will be enhanced at 2°C compared to 1.5°C global temperature rise (Ma et al. 2018). Adapting desertification control to climate change involves: improving the adaptation capacity to climate change for afforestation and grassland management by executing SLM practices; optimising the agricultural and animal husbandry structure; and using big data to meet the water resources regulation (Zhang and Huisingh 20181588). In particular, improving scientific and technological supports in desertification control is crucial for adaptation to climate change and combating desertification, including protecting vegetation in desertification-prone lands by planting indigenous plant species, facilitating natural restoration of vegetation to conserve biodiversity, employing artificial rain or snow, water-saving irrigation and water storage technologies (Jin et al. 2014; Yang et al. 20131589).

3.7.2.2

The Green Dam in Algeria

After independence in 1962, the Algerian government initiated measures to replant forests destroyed by the war, and the steppes affected by desertification, among its top priorities (Belaaz 20031591).

In 1972, the government invested in the Green Dam (Barrage vert) project. This was the first significant experiment to combat desertification, influence the local climate and decrease the aridity by restoring a barrier of trees. The Green Dam extends across arid and semi-arid zones between the isohyets 300 mm and 200 mm. It is a 3 Mha band of plantation running from east to west (Figure 3.12). It is over 1200 km long (from the Algerian–Moroccan border to the Algerian–Tunisian border) and has an average width of about 20 km. The soils in the area are shallow, low in organic matter and susceptible to erosion. The main objectives of the project were to conserve natural resources, improve the living conditions of local residents and avoid their exodus to urban areas. During the first four decades (1970–2000) the success rate was low (42%) due to lack of participation by the local population and the choice of species (Bensaid 19951592).

The Green Dam did not have the desired effects. Despite tree-planting efforts, desertification intensified on the steppes, especially in south-western Algeria, due to the prolonged drought during the 1980s. Rainfall declined in the range from 18% to 27%, and the dry season has increased by two months in the last century (Belala et al. 20181593). Livestock numbers in the Green Dam regions, mainly sheep, grew exponentially, leading to severe overgrazing, causing trampling and soil compaction, which greatly increased the risk of erosion. Wind erosion, very prevalent in the region, is due to climatic conditions and the strong anthropogenic action that reduced the vegetation cover. The action of the wind carries fine particles such as sands and clays and leaves on the soil surface a lag-gravel pavement, which is unproductive. Water erosion is largely due to torrential rains in the form of severe thunderstorms that disintegrate the bare soil surface from raindrop impact (Achite et al. 20161594). The detached soil and nutrients are transported offsite via runoff, resulting in loss of fertility and water holding capacity. The risk of and severity of water erosion is a function of human land-use activities that increase soil loss through removal of vegetative cover. The National Soil Sensitivity to Erosion Map (Salamani et al. 20121595) shows that more than 3 Mha of land in the steppe provinces are currently experiencing intense wind activity (Houyou et al. 20161596) and that these areas are at particular risk of soil erosion. Mostephaoui et al. (2013), estimates that each year there is a loss of 7 t ha–1 of soils due to erosion. Nearly 0.6 Mha of land in the steppe zone are fully degraded without the possibility of biological recovery.

To combat the effects of erosion and desertification, the government has planned to relaunch the rehabilitation of the Green Dam by incorporating new concepts related to sustainable development, and adaptation to climate change. The experience of previous years has led to integrated rangeland management, improved tree and fodder shrub plantations and the development of water conservation techniques. Reforestation is carried out using several species, including fruit trees, to increase and diversify the sources of income for the population.

The evaluation of the Green Dam from 1972 to 2015 (Merdas et al. 20151597) shows that 0.3 Mha of forest plantation have been planted, which represents 10% of the project area. Estimates of the success rate of reforestation vary considerably between 30% and 75%, depending on the region. Through demonstration, the Green Dam has inspired several African nations to work together to build a Great Green Wall to combat land degradation, mitigate climate change effects, loss of biodiversity and poverty in a region that stretches from Senegal to Djibouti (Sahara and Sahel Observatory (OSS) 2016) (Section 3.7.2.3).

Figure 3.12b

Location of the Green Dam in Algeria (Saifi et al. 2015). Note: The green coloured band represents the location of the Green Dam.

Location of the Green Dam in Algeria (Saifi et al. 20151806). Note: The green coloured band represents the location of the Green Dam.

3.7.2.3

The Great Green Wall of the Sahara and the Sahel Initiative

The Great Green Wall is an initiative of the Heads of State and Government of the Sahelo-Saharan countries to mitigate and adapt to climate change, and to improve the food security of the Sahel and Saharan peoples (Sacande 20181598; Mbow 20171599). Launched in 2007, this regional project aims to restore Africa’s degraded arid landscapes, reduce the loss of biodiversity and support local communities to sustainable use of forests and rangelands. The Great Green Wall focuses on establishing plantations and neighbouring projects, covering a distance of 7775 km from Senegal on the Atlantic coast to Eritrea on the Red Sea coast, with a width of 15 km (Figure 3.13). The wall passes through Djibouti, Eritrea, Ethiopia, Sudan, Chad, Niger, Nigeria, Mali, Burkina Faso, Mauritania and Senegal.

The choice of woody and herbaceous species that will be used to restore degraded ecosystems is based on biophysical and socio-economic criteria, including socio-economic value (food, pastoral, commercial, energetic, medicinal, cultural); ecological importance (carbon sequestration, soil cover, water infiltration); and resilience to climate change and variability. The Pan-African Agency of the Great Green Wall (PAGGW) was created in 2010 under the auspices of the African Union and CEN-SAD to manage the project. The initiative is implemented at the level of each country by a national structure. A monitoring and evaluation system has been defined, allowing nations to measure outcomes and to propose the necessary adjustments.

In the past, reforestation programmes in the arid regions of the Sahel and North Africa that have been undertaken to stop desertification were poorly studied and cost a lot of money without significant success (Benjaminsen and Hiernaux 20191600). Today, countries have changed their strategies and opted for rural development projects that can be more easily funded. Examples of scalable practices for land restoration include managing water bodies for livestock and crop production, and promoting fodder trees to reduce runoff (Mbow 20171601).

The implementation of the initiative has already started in several countries. For example, the FAO’s Action Against Desertification project was restoring 18,000 hectares of land in 2018 through planting native tree species in Burkina Faso, Ethiopia, The Gambia, Niger, Nigeria and Senegal (Sacande 20181602). Berrahmouni et al. (2016)1807 estimated that 166 Mha can be restored in the Sahel, requiring the restoration of 10 Mha per year to achieve Land Degradation Neutrality targets by 2030. Despite these early implementation actions on the ground, the achievement of the planned targets is questionable, and will be challenging without significant additional funding.

Figure 3.13

The Great Green Wall of the Sahara and the Sahel. Source for the data layer: This dataset is an extract from the GlobCover 2009 land cover map, covering Africa and the Arabian Peninsula. The GlobCover 2009 land cover map is derived by an automatic and regionally tuned classification of a time series of global MERIS […]

The Great Green Wall of the Sahara and the Sahel. Source for the data layer: This dataset is an extract from the GlobCover 2009 land cover map, covering Africa and the Arabian Peninsula. The GlobCover 2009 land cover map is derived by an automatic and regionally tuned classification of a time series of global MERIS (MEdium Resolution Imaging Spectrometer) FR mosaics for the year 2009. The global land cover map counts 22 land cover classes defined with the United Nations (UN) Land Cover Classification System (LCCS)

3.7.3

Invasive plant species

3.7.3.1

Introduction

The spread of invasive plants can be exacerbated by climate change (Bradley et al. 20101603; Davis et al. 20001604). In general, it is expected that the distribution of invasive plant species with high tolerance to drought or high temperatures may increase under most climate change scenarios (medium to high confidence) (Bradley et al. 20101605; Settele et al. 20141606; Scasta et al. 20151607). Invasive plants are considered a major risk to native biodiversity and can disturb the nutrient dynamics and water balance in affected ecosystems (Ehrenfeld 20031608). Compared to more humid regions, the number of species that succeed in invading dryland areas is low (Bradley et al. 20121609), yet they have a considerable impact on biodiversity and ecosystem services (Le Maitre et al. 2015, 2011; Newton et al. 20111610). Moreover, human activities in dryland areas are responsible for creating new invasion opportunities (Safriel et al. 20051611).

Current drivers of species introductions include expanding global trade and travel, land degradation and changes in climate (Chytrý et al. 20121612; Richardson et al. 20111613; Seebens et al. 20181614). For example, Davis et al. (2000) suggests that high rainfall variability promotes the success of alien plant species – as reported for semi-arid grasslands and Mediterranean-type ecosystems (Cassidy et al. 20041615; Reynolds et al. 20041616; Sala et al. 20061617). Furthermore, Panda et al. (2018) demonstrated that many invasive species could withstand elevated temperature and moisture scarcity caused by climate change. Dukes et al. (2011) observed that the invasive plant yellow-star thistle (Centaurea solstitialis) grew six time larger under the elevated atmospheric CO2 expected in future climate change scenarios.

Climate change is likely to aggravate the problem as existing species continue to spread unabated and other species develop invasive characteristics (Hellmann et al. 20081619). Although the effects of climate change on invasive species distributions have been relatively well explored, the greater impact on ecosystems is less well understood (Bradley et al. 20101620; Eldridge et al. 20111621).

Due to the time lag between the initial release of invasive species and their impact, the consequence of invasions is not immediately detected and may only be noticed centuries after introduction (Rouget et al. 20161622). Climate change and invading species may act in concert (Bellard et al. 20131623; Hellmann et al. 20081625; Seebens et al. 20151626). For example, invasion often changes the size and structure of fuel loads, which can lead to an increase in the frequency and intensity of fire (Evans et al. 2015). In areas where the climate is becoming warmer, an increase in the likelihood of suitable weather conditions for fire may promote invasive species, which in turn may lead to further desertification. Conversely, fire may promote plant invasions via several mechanisms (by reducing cover of competing vegetation, destroying native vegetation and clearing a path for invasive plants or creating favourable soil conditions) (Brooks et al. 20041627; Grace et al. 20011628; Keeley and Brennan 20121629).

Figure 3.14

Difference between the number of invasive alien species (n=99, from Bellard et al. (2013)) predicted to occur by 2050 (under A1B scenario) and current period ‘2000’ within the dryland areas

Difference between the number of invasive alien species (n=99, from Bellard et al. (2013)1808) predicted to occur by 2050 (under A1B scenario) and current period ‘2000’ within the dryland areas

At a regional scale, Bellard et al. (2013)1809 predicted increasing risk in Africa and Asia, with declining risk in Australia (Figure 3.14). This projection does not represent an exhaustive list of invasive alien species occurring in drylands.

A set of four case studies in Ethiopia, Mexico, the USA and Pakistan is presented below to describe the nuanced nature of invading plant species, their impact on drylands and their relationship with climate change.

3.7.3.2

Ethiopia

The two invasive plants that inflict the heaviest damage to ecosystems, especially biodiversity, are the annual herbaceous weed, Parthenium hysterophorus (Asteraceae) also known as Congress weed; and the tree species, Prosopis juliflora (Fabaceae) also called Mesquite, both originating from the southwestern United States to Central/South America (Adkins and Shabbir 20141630). Prosopis was introduced in the 1970s and has since spread rapidly. Prosopis, classified as the highest priority invader in Ethiopia, is threatening livestock production and challenging the sustainability of the pastoral systems. Parthenium is believed to have been introduced along with relief aid during the debilitating droughts of the early 1980s, and a recent study reported that it has spread into 32 out of 34 districts in Tigray, the northernmost region of Ethiopia (Teka 20161631). A study by Etana et al. (2011) indicated that Parthenium caused a 69% decline in the density of herbaceous species in Awash National Park within a few years of introduction. In the presence of Parthenium, the growth and development of crops is suppressed due to its allelopathic properties. McConnachie et al. (2011) estimated a 28% crop loss across the country, including a 40–90% reduction in sorghum yield in eastern Ethiopia alone (Tamado et al. 20021632). The weed is a substantial agricultural and natural resource problem and constitutes a significant health hazard (Fasil 2011). Parthenium causes acute allergic respiratory problems, skin dermatitis, and reportedly mutagenicity both in humans and livestock (Mekonnen 2017; Patel 20111633). The eastern belt of Africa – including Ethiopia – presents a very suitable habitat, and the weed is expected to spread further in the region in the future (Mainali et al. 20151635).

There is neither a comprehensive intervention plan nor a clear institutional mandate to deal with invasive weeds, however, there are fragmented efforts involving local communities even though they are clearly inadequate. The lessons learned, related to actions that have contributed to the current scenario, are several. First, lack of coordination and awareness – mesquite was introduced by development agencies as a drought-tolerant shade tree with little consideration of its invasive nature. If research and development institutions had been aware, a containment strategy could have been implemented early on. The second major lesson is the cost of inaction. When research and development organisations did sound the alarm, the warnings went largely unheeded, resulting in the spread and buildup of two of the worst invasive plant species in the world (Fasil 20111636).

3.7.3.3

Mexico

Buffelgrass (Cenchrus ciliaris L.), a native species from southern Asia and East Africa, was introduced into Texas and northern Mexico in the 1930s and 1940s, as it is highly productive in drought conditions (Cox et al. 1988; Rao et al. 1996). In the Sonoran desert of Mexico, the distribution of buffelgrass has increased exponentially, covering 1 Mha in Sonora State (Castellanos-Villegas et al. 20021637). Furthermore, its potential distribution extended to 53% of Sonora State and 12% of semi-arid and arid ecosystems in Mexico (Arriaga et al. 20041638). Buffelgrass has also been reported as an aggressive invader in Australia and the USA, resulting in altered fire cycles that enhance further spread of this plant and disrupt ecosystem processes (Marshall et al. 20121639; Miller et al. 20101641; Schlesinger et al. 20131642).

Castellanos et al. (2016) reported that soil moisture was lower in the buffelgrass savannah cleared 35 years ago than in the native semi-arid shrubland, mainly during the summer. The ecohydrological changes induced by buffelgrass can therefore displace native plant species over the long term. Invasion by buffelgrass can also affect landscape productivity, as it is not as productive as native vegetation (Franklin and Molina-Freaner 20101643). Incorporation of buffelgrass is considered a good management practice by producers and the government. For this reason, no remedial actions are undertaken.

3.7.3.4

United States of America

Sagebrush ecosystems have declined from 25 Mha to 13 Mha since the late 1800s (Miller et al. 20111644). A major cause is the introduction of non-native cheatgrass (Bromus tectorum), which is the most prolific invasive plant in the USA. Cheatgrass infests more than 10 Mha in the Great Basin and is expanding every year (Balch et al. 20131645). It provides a fine-textured fuel that increases the intensity, frequency and spatial extent of fire (Balch et al. 2013). Historically, wildfire frequency was 60 to 110 years in Wyoming big sagebrush communities and has increased to five years following the introduction of cheatgrass (Balch et al. 20131646; Pilliod et al. 20171648).

The conversion of the sagebrush steppe biome to annual grassland with higher fire frequencies has severely impacted livestock producers, as grazing is not possible for a minimum of two years after fire. Furthermore, cheatgrass and wildfires reduce critical habitat for wildlife and negatively impact species richness and abundance – for example, the greater sage-grouse (Centocercus urophasianus) and pygmy rabbit (Brachylagus idahoensis) which are on the verge of being listed for federal protection (Crawford et al. 20041649; Larrucea and Brussard 20081650; Lockyer et al. 20151651).

Attempts to reduce cheatgrass impacts through reseeding of both native and adapted introduced species have occurred for more than 60 years (Hull and Stewart 19491652) with little success. Following fire, cheatgrass becomes dominant and recovery of native shrubs and grasses is improbable, particularly in relatively low-elevation sites with minimal annual precipitation (less than 200 mm yr–1) (Davies et al. 20121653; Taylor et al. 20141654). Current rehabilitation efforts emphasise the use of native and non-native perennial grasses, forbs and shrubs (Bureau of Land Management 20051655). Recent literature suggests that these treatments are not consistently effective at displacing cheatgrass populations or re-establishing sage-grouse habitat, with success varying with elevation and precipitation (Arkle et al. 20141656; Knutson et al. 20141657). Proper post-fire grazing rest, season-of-use, stocking rates, and subsequent management are essential to restore resilient sagebrush ecosystems before they cross a threshold and become an annual grassland (Chambers et al. 20141658; Miller et al. 20111659; Pellant et al. 20041660). Biological soil crust protection may be an effective measure to reduce cheatgrass germination, as biocrust disturbance has been shown to be a key factor promoting germination of non-native grasses (Hernandez and Sandquist 2011). Projections of increasing temperature (Abatzoglou and Kolden 20111662), and observed reductions in and earlier melting of snowpack in the Great Basin region (Harpold and Brooks 20181663; Mote et al. 20051664) suggest that there is a need to understand current and past climatic variability as this will drive wildfire variability and invasions of annual grasses.

3.7.3.5

Pakistan

The alien plants invading local vegetation in Pakistan include Brossentia papyrifera (found in Islamabad Capital territory), Parthenium hysterophorus (found in Punjab and Khyber Pakhtunkhwa provinces), Prosopis juliflora (found all over Pakistan), Eucalyptus camaldulensis (found in Punjab and Sindh provinces), Salvinia (aquatic plant widely distributed in water bodies in Sindh), Cannabis sativa (found in Islamabad Capital Territory), Lantana camara and Xanthium strumarium (found in upper Punjab and Khyber Pakhtunkhwa provinces) (Khan et al. 20101665; Qureshi et al. 20141666). Most of these plants were introduced by the Forest Department decades ago for filling the gap between demand and supply of timber, fuelwood and fodder. These non-native plants have some uses but their disadvantages outweigh their benefits (Marwat et al. 20101667; Rashid et al. 20141668).

Besides being a source of biological pollution and a threat to biodiversity and habitat loss, the alien plants reduce the land value and cause huge losses to agricultural communities (Rashid et al. 20141810). Brossentia papyrifera, commonly known as Paper Mulberry, is the root cause of inhalant pollen allergy for the residents of lush green Islamabad during spring. From February to April, the pollen allergy is at its peak, with symptoms of severe persistent coughing, difficulty in breathing, and wheezing. The pollen count, although variable at different times and days, can be as high as 55,000 m-3.

Early symptoms of the allergy include sneezing, itching in the eyes and skin, and blocked nose. With changing climate, the onset of disease is getting earlier, and pollen count is estimated to cross 55,000 m–3 (Rashid et al. 20141670). About 45% of allergic patients in the twin cities of Islamabad and Rawalpindi showed positive sensitivity to the pollens (Marwat et al. 20101671). Millions of rupees have been spent by the Capital Development Authority on pruning and cutting of Paper Mulberry trees but because of its regeneration capacity growth is regained rapidly (Rashid et al. 20141672). Among other invading plants, Prosopis juliflora has allelopathic properties, and Eucalyptus is known to transpire huge amounts of water and deplete the soil of its nutrient elements (Qureshi et al. 20141673).

Although a Biodiversity Action Plan exists in Pakistan, it is not implemented in letter or spirit. The Quarantine Department focuses only on pests and pathogens but takes no notice of plant and animal species being imported. Also, there is no provision for checking the possible impacts of imported species on the environment (Rashid et al. 20141674) or for carrying out bioassays of active allelopathic compounds of alien plants.

3.7.4

Oases in hyper-arid areas in the Arabian Peninsula and northern Africa

Oases are isolated areas with reliable water supply from lakes and springs, located in hyper-arid and arid zones (Figure 3.15). Oasis agriculture has long been the only viable crop production system throughout the hot and arid regions of the Arabian Peninsula and North Africa. Oases in hyper-arid climates are usually subject to water shortage as evapotranspiration exceeds rainfall. This often causes salinisation of soils. While many oases have persisted for several thousand years, many others have been abandoned, often in response to changes in climate or hydrologic conditions (Jones et al. 20191675), providing testimony to societies’ vulnerability to climatic shifts and raising concerns about similarly severe effects of anthropogenic climate change (Jones et al. 20191676).

On the Arabian Peninsula and in North Africa, climate change is projected to have substantial and complex effects on oasis areas (Abatzoglou and Kolden 20111677; Ashkenazy et al. 20121678; Bachelet et al. 20161679; Guan et al. 20181680; Iknayan and Beissinger 20181681; Ling et al. 20131682). To illustrate, by the 2050s, the oases in southern Tunisia are expected to be affected by hydrological and thermal changes, with an average temperature increase of 2.7°C, a 29% decrease in precipitation and a 14% increase in evapotranspiration rate (Ministry of Agriculture and Water Resources of Tunisia and GIZ 20071683). In Morocco, declining aquifer recharge is expected to impact the water supply of the Figuig oasis (Jilali 20141684), as well as for the Draa Valley (Karmaoui et al. 20161685). Saudi Arabia is expected to experience a 1.8°C–4.1°C increase in temperatures by 2050, which is forecast to raise agricultural water demand by 5–15% in order to maintain production levels equal to those of 2011 (Chowdhury and Al-Zahrani 20131686). The increase of temperatures and variable pattern of rainfall over the central, north and south-western regions of Saudi Arabia may pose challenges for sustainable water resource management (Tarawneh and Chowdhury 20181687). Moreover, future climate scenarios are expected to increase the frequency of floods and flash floods, such as in the coastal areas along the central parts of the Red Sea and the south-southwestern areas of Saudi Arabia (Almazroui et al. 20171688).

While many oases are cultivated with very heat-tolerant crops such as date palms, even such crops eventually have declines in their productivity when temperatures exceed certain thresholds or hot conditions prevail for extended periods. Projections so far do not indicate severe losses in land suitability for date palm for the Arabian Peninsula (Aldababseh et al. 20181689; Shabani et al. 20151690). It is unclear, however, how reliable the climate response parameters in the underlying models are, and actual responses may differ substantially.

Figure 3.15a

Oases across the Arabian Peninsula and North Africa (alphabetically by country). (a) Masayrat ar Ruwajah oasis, Ad Dakhiliyah Governorate, Oman (Photo: Eike Lüdeling).

Oases across the Arabian Peninsula and North Africa (alphabetically by country). (a) Masayrat ar Ruwajah oasis, Ad Dakhiliyah Governorate, Oman (Photo: Eike Lüdeling).

Figure 3.15b

(b) Tasselmanet oasis, Ouarzazate Province, Morocco (Photo: Abdellatif Khattabi).

(b) Tasselmanet oasis, Ouarzazate Province, Morocco (Photo: Abdellatif Khattabi).

Figure 3.15c

c) Al-Ahsa oasis, Al-Ahsa Governarate, Saudi Arabia (Photo: Shijan Kaakkara).

c) Al-Ahsa oasis, Al-Ahsa Governarate, Saudi Arabia (Photo: Shijan Kaakkara).

Figure-3.15d

Zarat oasis, Governorate of Gabes, Tunisia (Photo: Hamda Aloui). The use rights for (a), (b) and (d) were granted by copyright holders; (c) is licensed under the Creative Commons Attribution 2.0 Generic license.

Zarat oasis, Governorate of Gabes, Tunisia (Photo: Hamda Aloui). The use rights for (a), (b) and (d) were granted by copyright holders; (c) is licensed under the Creative Commons Attribution 2.0 Generic license.

Date palms are routinely assumed to be able to endure very high temperatures, but recent transcriptomic and metabolomic evidence suggests that heat stress reactions already occur at 35°C (Safronov et al. 20171691), which is not exceptionally warm for many oases in the region. Given current assumptions about the heat-tolerance of date palm, however, adverse effects are expected to be small (Aldababseh et al. 20181692; Shabani et al. 20151693). For some other perennial oasis crops, impacts of temperature increases are already apparent. Between 2004/2005 and 2012/2013, high-mountain oases of Al Jabal Al Akhdar in Oman lost almost all fruit and nut trees of temperate-zone origin, with the abundance of peaches, apricots, grapes, figs, pears, apples, and plums dropping by between 86% and 100% (Al-Kalbani et al. 20161694). This implies that that the local climate may not remain suitable for species that depend on cool winters to break their dormancy period (Luedeling et al. 20091695). A similar impact is very probable in Tunisia and Morocco, as well as in other oasis locations in the Arabian Peninsula and North Africa (Benmoussa et al. 20071811). All these studies expect strong decreases in winter chill, raising concerns that many currently well-established species will no longer be viable in locations where they are grown today. The risk of detrimental chill shortfalls is expected to increase gradually, slowly diminishing the economic prospects to produce such species. Without adequate adaptation actions, the consequences of this development for many traditional oasis settlements and other plantations of similar species could be highly negative.

At the same time, population growth and agricultural expansion in many oasis settlements are leading to substantial increases in water demand for human consumption (Al-Kalbani et al. 20141696). For example, a large unmet water demand has been projected for future scenarios in the valley of Seybouse in East Algeria (Aoun-Sebaiti et al. 20141697), and similar conclusions were drawn for Wadi El Natrun in Egypt (Switzman et al. 20181698). Modelling studies have indicated long-term decline in available water and increasing risk of water shortages – for example, for oases in Morocco (Johannsen et al. 20161699; Karmaoui et al. 20161700), the Dakhla oasis in Egypt’s Western Desert (Sefelnasr et al. 20141701) and for the large Upper Mega Aquifer of the Arabian Peninsula (Siebert et al. 20161702). Mainly due to the risk of water shortages, Souissi et al. (2018) classified almost half of all farmers in Tunisia as non-resilient to climate change, especially those relying on tree crops, which limit opportunities for short-term adaptation actions.

The maintenance of the oasis systems and the safeguarding of their population’s livelihoods are currently threatened by continuous water degradation, increasing soil salinisation, and soil contamination (Besser et al. 20171703). Waterlogging and salinisation of soils due to rising saline groundwater tables coupled with inefficient drainage systems have become common to all continental oases in Tunisia, most of which are concentrated around saline depressions, known locally as chotts (Ben Hassine et al. 20131704). Similar processes of salinisation are also occurring in the oasis areas of Egypt due to agricultural expansion, excessive use of water for irrigation and deficiency of the drainage systems (Abo-Ragab 20101705; Masoud and Koike 20061706). A prime example for this is Siwa oasis (Figure 3.16), a depression extending over 1050 km2 in the north-western desert of Egypt in the north of the sand dune belt of the Great Sand Sea (Abo-Ragab and Zaghloul 20171707). Siwa oasis has been recognised as a Globally Important Agricultural Heritage Site (GIAHS) by the FAO for being an in situ repository of plant genetic resources, especially of uniquely adapted varieties of date palm, olive and secondary crops that are highly esteemed for their quality and continue to play a significant role in rural livelihoods and diets (FAO 2016).

The population growth in Siwa is leading rapid agricultural expansion and land reclamation.The Siwan farmers are converting the surrounding desert into reclaimed land by applying their old inherited traditional practices. Yet, agricultural expansion in the oasis mainly depends on non-renewable groundwaters. Soil salinisation and vegetation loss have been accelerating since 2000 due to water mismanagement and improper drainage systems (Masoud and Koike 20061708). Between 1990 and 2008, the cultivated area increased from 53 to 88 km , lakes from 60 to 76 km2 , sabkhas (salt flats) from 335 to 470 km2, and the urban area from 6 to 10 km2 (Abo-Ragab 20101709). The problem of rising groundwater tables was exacerbated by climatic changes (Askri et al. 20101710; Gad and Abdel-Baki 2002; Marlet et al. 20091711).

Water supply is likely to become even scarcer for oasis agriculture under changing climate in the future than it is today, and viable solutions are difficult to find. While some authors stress the possibility to use desalinated water for irrigation (Aldababseh et al. 20181712), the economics of such options, especially given the high evapotranspiration rates in the Arabian Peninsula and North Africa, are debatable. Many oases are located far from water sources that are suitable for desalination, adding further to feasibility constraints. Most authors therefore stress the need to limit water use (Sefelnasr et al. 20141713), for example, by raising irrigation efficiency (Switzman et al. 20181714), reducing agricultural areas (Johannsen et al. 20161715) or imposing water use restrictions (Odhiambo 20171716), and to carefully monitor desertification (King and Thomas 20141717). Whether adoption of crops with low water demand, such as sorghum (Sorghum bicolor (L.) Moench) or jojoba (Simmondsia chinensis (Link) C. K. Schneid.) (Aldababseh et al. 20181718), can be a viable option for some oases remains to be seen, but given their relatively low profit margins compared to currently grown oasis crops, there are reasons to doubt the economic feasibility of such proposals. While it is currently unclear to what extent oasis agriculture can be maintained in hot locations of the region, cooler sites offer potential for shifting towards new species and cultivars, especially for tree crops, which have particular climatic needs across seasons. Resilient options can be identified, but procedures to match tree species and cultivars with site climate need to be improved to facilitate effective adaptation.

There is high confidence that many oases of North Africa and the Arabian Peninsula are vulnerable to climate change. While the impacts of recent climate change are difficult to separate from the consequences of other change processes, it is likely that water resources have already declined in many places and the suitability of the local climate for many crops, especially perennial crops, has already decreased. This decline of water resources and thermal suitability of oasis locations for traditional crops is very likely to continue throughout the 21st century. In the coming years, the people living in oasis regions across the world will face challenges due to increasing impacts of global environmental change (Chen et al. 20181719). Hence, efforts to increase their adaptive capacity to climate change can facilitate the sustainable development of oasis regions globally. In particular this will mean addressing the trade-offs between environmental restoration and agricultural livelihoods (Chen et al. 2018). Ultimately, sustainability in oasis regions will depend on policies integrating the provision of ecosystem services and social and human welfare needs (Wang et al. 20171724).

Figure 3.16

Satellite image of the Siwa Oasis, Egypt. Source: Google Maps

Satellite image of the Siwa Oasis, Egypt. Source: Google Maps

3.7.5

Integrated watershed management

Desertification has resulted in significant loss of ecosystem processes and services, as described in detail in this chapter. The techniques and processes to restore degraded watersheds are not linear and integrated watershed management (IWM) must address physical, biological and social approaches to achieve SLM objectives (German et al. 20071726).

3.7.5.1

Jordan

Population growth, migration into Jordan and changes in climate have resulted in desertification of the Jordan Badia region. The Badia region covers more than 80% of the country’s area and receives less than 200 mm of rainfall per year, with some areas receiving less than 100 mm (Al-Tabini et al. 20121727). Climate analysis has indicated a generally increasing dryness over the West Asia and Middle East region (AlSarmi and Washington 20111728; Tanarhte et al. 20151729), with reduction in average annual rainfall in Jordan’s Badia area (De Pauw et al. 20151730). The incidence of extreme rainfall events has not declined over the region. Locally increased incidence of extreme events over the Mediterranean region has been proposed (Giannakopoulos et al. 20091731).

The practice of intensive and localised livestock herding, in combination with deep ploughing and unproductive barley agriculture, are the main drivers of severe land degradation and depletion of the rangeland natural resources. This affected both the quantity and the diversity of vegetation as native plants with a high nutrition value were replaced with invasive species with low palatability and nutritional content (Abu-Zanat et al. 20041732). The sparsely covered and crusted soils in Jordan’s Badia area have a low rainfall interception and infiltration rate, which leads to increased surface runoff and subsequent erosion and gullying, speeding up the drainage of rainwater from the watersheds, which can result in downstream flooding in Amman, Jordan (Oweis 20171733).

Figure 3.17a

(a)Newly prepared micro water harvesting catchment, using the Vallerani system.

(a)Newly prepared micro water harvesting catchment, using the Vallerani system.

Figure-3.17b

(b) Aerial imaging showing micro water harvesting catchment treatment after planting

(b) Aerial imaging showing micro water harvesting catchment treatment after planting

Figure 3.17c

(c) one year after treatment. Source: Stefan Strohmeier.

(c) one year after treatment. Source: Stefan Strohmeier.

Figure 3.18

Illustration of enhanced soil water retention in the Mechanized Micro Rainwater Harvesting compared to untreated Badia rangelands in Jordan, showing precipitation (PCP), sustained stress level resulting in decreased production, field capacity and wilting point for available soil moisture, and then measured soil moisture content between the two treatments (degraded rangeland and the restored rangeland with […]

Illustration of enhanced soil water retention in the Mechanized Micro Rainwater Harvesting compared to untreated Badia rangelands in Jordan, showing precipitation (PCP), sustained stress level resulting in decreased production, field capacity and wilting point for available soil moisture, and then measured soil moisture content between the two treatments (degraded rangeland and the restored rangeland with the Vallerani plough).

To restore the desertified Badia an IWM plan was developed using hillslope-implemented water harvesting micro catchments as a targeted restoration approach (Tabieh et al. 20151734). Mechanized Micro Rainwater Harvesting (MIRWH) technology using the ‘Vallerani plough’ (Antinori and Vallerani 19941735; Gammoh and Oweis 20111736; Ngigi 20031737) is being widely applied for rehabilitation of highly degraded rangeland areas in Jordan. A tractor digs out small water harvesting pits on the contour of the slope (Figure 3.17) allowing the retention, infiltration and local storage of surface runoff in the soil (Oweis 20171739). The micro catchments are planted with native shrub seedlings, such as saltbush (Atriplex halimus), with enhanced survival as a function of increased soil moisture (Figure 3.18) and increased dry matter yields (>300 kg ha–1) that can serve as forage for livestock (Oweis 20171738; Tabieh et al. 20151740).

Simultaneously to MIRWH upland measures, the gully erosion is being treated through intermittent stone plug intervention (Figure 3.19), stabilising the gully beds, increasing soil moisture in proximity of the plugs, dissipating the surface runoff’s energy, and mitigating further back-cutting erosion and quick drainage of water. Eventually, the treated gully areas silt up and dense vegetation cover can re-establish. In addition, grazing management practices are implemented to increase the longevity of the treatment. Ultimately, the recruitment processes and re-vegetation shall control the watershed’s hydrological regime through rainfall interception, surface runoff deceleration and filtration, combined with the less erodible and enhanced infiltration characteristics of the rehabilitated soils.

Figure 3.19a

(a) Gully plug development in September 2017.

(a) Gully plug development in September 2017.

Figure 3.19b

(b) Post-rainfall event (March 2018). Near Amman, Jordan. Source: Stefan Strohmeier.

(b) Post-rainfall event (March 2018). Near Amman, Jordan. Source: Stefan Strohmeier.

In-depth understanding of the Badia’s rangeland status transition, coupled with sustainable rangeland management, are still subject to further investigation, development and adoption; a combination of all three is required to mitigate the ongoing degradation of the Middle Eastern rangeland ecosystems.

Oweis (2017)1813 indicated that the cost of the fully automated Vallerani technique was approximately 32 USD ha-1. The total cost of the restoration package included the production, planting and maintenance of the shrub seedlings (11 USD ha–1). Tabieh et al. (2015)1812 calculated a benefit-cost ratio (BCR) of above 1.5 for re-vegetation of degraded Badia areas through MIRWH and saltbush. However, costs vary based on the seedling’s costs and availability of trained labour.

Water harvesting is not a recent scientific advancement. Water harvesting is known to have been developed during the Bronze Age and was widely practiced in the Negev Desert during the Byzantine time period (1300–1600 years ago) (Fried et al. 20181741; Stavi et al. 20171742). Through construction of various structures made of packed clay and stone, water was either held on site in half-circular dam structures (hafir) that faced up-slope to capture runoff, or on terraces that slowed water allowing it to infiltrate and to be stored in the soil profile. Numerous other systems were designed to capture water in below-ground cisterns to be used later to provide water to livestock or for domestic use. Other water harvesting techniques divert runoff from hillslopes or wadis and spread the water in a systematic manner across playas and the toe-slope of a hillslope. These systems allow production of crops in areas with 100 mm of average annual precipitation by harvesting an additional 300+ mm of water (Beckers et al. 20131743). Water harvesting is a proven technology to mitigate or adapt to climate change where precipitation may be reduced, and allow for small-scale crop and livestock production to continue supporting local needs.

3.7.5.2

India

The second great challenge after the Green Revolution in India was the low productivity in the rain-fed and semi-arid regions where land degredation and drought were serious concerns. In response to this challenge IWM projects were implemented over large areas in semi-arid biomes over the past few decades. IWM was meant to become a key factor in meeting a range of social development goals in many semi-arid rainfed agrarian landscapes in India (Bouma et al. 20071744; Kerr et al. 20021745). Over the years, watershed development has become the fulcrum of rural development, and has the potential to achieve the twin objectives of ecosystem restoration and livelihood assurance in the drylands of India (Joy et al. 2004).

Many reports indicate significant improvements in mitigation of drought impacts, raising crops and fodder, livestock productivity, expanding the availability of drinking water and increasing incomes as a result of IWM (Rao 2000), but in some cases overall the positive impact of the programme has been questioned and, except in a few cases, the performance has not lived up to expectations (Joy et al. 2004; JM Kerr et al. 2002). Comparisons of catchments with and without IWM projects using remotely sensed data have sometimes shown no significant enhancement of biomass, in part due to methodological challenges of space for time comparisons (Bhalla et al. 20131746). The factors contributing to the successful cases were found to include effective participation of stakeholders in management (Rao 2000; Ratna Reddy et al. 20041747).

Attribution of success in soil and water conservation measures was confounded by inadequate monitoring of rainfall variability and lack of catchment hydrologic indicators (Bhalla et al. 20131748). Social and economic trade-offs included bias of benefits to downstream crop producers at the expense of pastoralists, women and upstream communities. This biased distribution of IWM benefits could potentially be addressed by compensation for environmental services between communities (Kerr et al. 20021749). The successes in some areas also led to increased demand for water, especially groundwater, since there has been no corresponding social regulation of water use after improvement in water regime (Samuel et al. 20071750). Policies and management did not ensure water allocation to sectors with the highest social and economic benefits (Batchelor et al. 20031751). Limited field evidence of the positive impacts of rainwater harvesting at the local scale is available, but there are several potential negative impacts at the watershed scale (Glendenning et al. 20121752). Furthermore, watershed projects are known to have led to more water scarcity, and higher expectations for irrigation water supply, further exacerbating water scarcity (Bharucha et al. 20141753).

In summary, the mixed performance of IWM projects has been linked to several factors. These include: inequity in the distribution of benefits (Kerr et al. 2002); focus on institutional aspects rather than application of appropriate watershed techniques and functional aspects of watershed restoration (Joy et al. 2006; Vaidyanathan 20061755); mismatch between scales of focus and those that are optimal for catchment processes (Kerr 20071756); inconsistencies in criteria used to select watersheds for IWM projects (Bhalla et al. 20111757); and in a few cases additional costs and inefficiencies of local non-governmental organisations (Chandrasekhar et al. 20061758; Deshpande 20081759). Enabling policy responses for improvement of IWM performance include: a greater emphasis on ecological restoration rather than civil engineering; sharper focus on sustainability of livelihoods than just conservation; adoption of ‘water justice’ as a normative goal and minimising externalities on non-stakeholder communities; rigorous independent biophysical monitoring, with feedback mechanisms and integration with larger schemes for food and ecological security, and maintenance of environmental flows for downstream areas (Bharucha et al. 20141760; Calder et al. 20081761; Joy et al. 2006). Successful adaptation of IWM to achieve land degradation neutrality would largely depend on how IWM creatively engages with dynamics of large-scale land use and hydrology under a changing climate, involvement of livelihoods and rural incomes in ecological restoration, regulation of groundwater use, and changing aspirations of rural population (robust evidence, high agreement) (O’Brien et al. 20041762; Samuel et al. 20071763; Samuel and Joy 20181764).

3.7.5.3

Limpopo River Basin

Covering an area of 412,938 km2, the Limpopo River basin spans parts of Botswana, South Africa, Zimbabwe and Mozambique, eventually entering into the Mozambique Channel. It has been selected as a case study as it provides a clear illustration of the combined effect of desertification and climate change, and why IWM may be a crucial component of reducing exposure to climate change. It is predominantly a semi-arid area with an average annual rainfall of 400 mm (Mosase and Ahiablame 20181765). Rainfall is both highly seasonal and variable, with the prominent impact of the El Niño/ La Niña phenomena and the Southern Oscillation leading to severe droughts (Jury 20161766). It is also exposed to tropical cyclones that sweep in from the Mozambique Channel often leading to extensive casualties and the destruction of infrastructure (Christie and Hanlon 20011767). Furthermore, there is good agreement across climate models that the region is going to become warmer and drier, with a change in the frequency of floods and droughts (Engelbrecht et al. 20111768; Zhu and Ringler 2012). Seasonality is predicted to increase, which in turn may increase the frequency of flood events in an area that is already susceptible to flooding (Spaliviero et al. 20141769).

A clear need exists to both address exposure to flood events as well as predicted decreases in water availability, which are already acute. Without the additional impact of climate change, the basin is rapidly reaching a point where all available water has been allocated to users (Kahinda et al. 20161770; Zhu and Ringler 2012). The urgency of the situation was identified several decades ago (FAO 2004), with the countries of the basin recognising that responses are required at several levels, both in terms of system governance and the need to address land degradation.

Recent reviews of the governance and implementation of IWM within the basin recognise that an integrated approach is needed and that a robust institutional, legal, political, operational, technical and support environment is crucial (Alba et al. 20161771; Gbetibouo et al. 20101773; Machethe et al. 20041774; Spaliviero et al. 20111775; van der Zaag and Savenije 19991776). Within the scope of emerging lessons, two principal ones emerge. The first is capacity and resource constraints at most levels. Limited capacity within Limpopo Watercourse Commission (LIMCOM) and national water management authorities constrains the implementation of IWM planning processes (Kahinda et al. 20161777; Spaliviero et al. 20111778). Whereas strategy development is often relatively well-funded and resourced through donor funding, long-term implementation is often limited due to competing priorities. The second is adequate representation of all parties in the process in order to address existing inequalities and ensure full integration of water management. For example, within Mozambique, significant strides have been made towards the decentralisation of river basin governance and IWM. Despite good progress, Alba et al. (2016) found that the newly implemented system may enforce existing inequalities as not all stakeholders, particularly smallholder farmers, are adequately represented in emerging water management structures and are often inhibited by financial and institutional constraints. Recognising economic and socio-political inequalities, and explicitly considering them to ensure the representation of all participants, can increase the chances of successful IWM implementation.

3.8

Knowledge gaps and key uncertainties

  • Desertification has been studied for decades and different drivers of desertification have been described, classified, and are generally understood (e.g., overgrazing by livestock or salinisation from inappropriate irrigation) (D’Odorico et al. 20131779). However, there are knowledge gaps on the extent and severity of desertification at global, regional, and local scales (Zhang and Huisingh 20181780; Zucca et al. 20121781). Overall, improved estimation and mapping of areas undergoing desertification is needed. This requires a combination of rapidly expanding sources of remotely sensed data, ground observations and new modelling approaches. This is a critical gap, especially in the context of measuring progress towards achieving the Land Degradation Neutrality target by 2030 in the framework of SDGs.
  • Despite numerous relevant studies, consistent indicators for attributing desertification to climatic and/or human causes are still lacking due to methodological shortcomings.
  • Climate change impacts on dust and sand storm activity remain a critical gap. In addition, the impacts of dust and sand storms on human welfare, ecosystems, crop productivity and animal health are not measured, particularly in the highly affected regions such as the Sahel, North Africa, the Middle East and Central Asia. Dust deposition on snow and ice has been found in many regions of the globe (e.g., Painter et al. 2018; Kaspari et al. 20141782; Qian et al. 20151783; Painter et al. 20131784), however, the quantification of the effect globally, and estimation of future changes in the extent of this effect, remain knowledge gaps.
  • Future projections of combined impacts of desertification and climate change on ecosystem services, fauna and flora, are lacking, even though this topic is of considerable social importance. Available information is mostly on separate, individual impacts of either (mostly) climate change or desertification. Responses to desertification are species-specific and mechanistic models are not yet able to accurately predict individual species responses to the many factors associated with desertification under changing climate.
  • Previous studies have focused on the general characteristics of past and current desertification feedbacks to the climate system. However, the information on the future interactions between climate and desertification (beyond changes in the aridity index) are lacking. The knowledge of future climate change impacts on such desertification processes as soil erosion, salinisation, and nutrient depletion remains limited both at the global and at the local levels.
  • Further research to develop the technologies and innovations needed to combat desertification is required, but it is also important to gain a better understanding of the reasons for the observed poor adoption of available innovations, to improve adoption rates.
  • Desertification under changing climate has a high potential to increase poverty, particularly through the risks coming from extreme weather events (Olsson et al. 20141785). However, the evidence rigorously attributing changes in observed poverty to climate change impacts is currently not available.
  • The knowledge on the limits to adaptation to the combined effects of climate change and desertification is insufficient. This is an important gap since the potential for residual risks and maladaptive outcomes is high.
  • Filling these gaps involves considerable investments in research and data collection. Using Earth observation systems in a standardised approach could help fill some of these gaps. This would increase data comparability and reduce uncertainty in approaches and costs. Systematically collected data would provide far greater insights than incomparable fragmented data.

Footnotes

  1. Pastoralists derive more than 50% of their income from livestock and livestock products, whereas agropastoralists generate more than 50% of their income from crop production and at least 25% from livestock production (Swift,1988).
  2. The estimates of the number of pastoralists, and especially of nomadic pastoralists, are very uncertain, because often nomadic pastoralists are not fully captured in national surveys and censuses (Carr-Hill, 2013).

References

  1. UNEP, 1992: World Atlas of Desertification [Middleton, N. and D.S.G. Thomas (eds.)]. UNEP.Edward Arnold, London, UK, 69 pp.
  2. van der Esch, S. et al., 2017: Exploring Future Changes in Land Use and Land Condition and the Impacts on Food, Water, Climate Change and Biodiversity: Scenarios for the UNCCD Global Land Outlook. Policy Report, PBL Netherlands Environmental Assessment Agency, the Hague, Netherlands, 115 pp.
  3. Koutroulis, A.G., 2019: Dryland changes under different levels of global warming. Sci. Total Environ., 655, 482–511, doi:10.1016/J.SCITOTENV.2018.11.215.
  4. Prăvălie, R., 2016: Drylands extent and environmental issues. A global approach. Earth-Science Rev., 161, 259–278, doi:10.1016/J.EARSCIREV.2016.08.003.
  5. UNCCD, 1994: Elaboration of an international convention to combat desertification in countries experiencing serious drought and/or desertification, particularly in Africa. General Assembly, United Nations, 1–58 pp.
  6. UNEP, 2006: Global Deserts Outlook. United Nations Environment Programme, Nairobi, Kenya, 148 pp.
  7. Safriel, U., 2009: Deserts and desertification: Challenges but also opportunities. L. Degrad. Dev., 20, 353–366, doi:10.1002/ldr.935.
  8. Gbeckor-Kove, N., 1989: Drought and Desertification. World Meteorological Organization, Geneva, Switzerland, 41–73, 286 pp.
  9. Türkeş. M., 1999: Vulnerability of Turkey to desertification with respect to precipitation and aridity conditions. Turkish J. Eng. Environ. Sci., 23, 363–380.
  10. Maliva, R., and T. Missimer, 2012: Aridity and drought. In: Arid Lands Water Evaluation and Management [Maliva, R. and T. Missimer (eds.)]. Springer, Berlin, Germany, pp. 21–39.
  11. Wilhite, D.A., M.V.K. Sivakumar, and R. Pulwarty, 2014: Managing drought risk in a changing climate: The role of national drought policy. Weather Clim. Extrem., 3, 4–13, doi:10.1016/J.WACE.2014.01.002.
  12. Sheffield, J., E.F. Wood, and M.L. Roderick, 2012: Little change in global drought over the past 60 years. Nature, 491, 435–438, doi:10.1038/ nature11575.
  13. Sheffield, J., E.F. Wood, and M.L. Roderick, 2012: Little change in global drought over the past 60 years. Nature, 491, 435–438, doi:10.1038/nature11575.
  14. Greve, P., B. Orlowsky, B. Mueller, J. Sheffield, M. Reichstein, and 
S.I. Seneviratne, 2014: Global assessment of trends in wetting and drying over land. Nat. Geosci., 7, 716–721, doi:10.1038/ngeo2247.
  15. Roderick, M.L., P. Greve, and G.D. Farquhar, 2015: On the assessment of aridity with changes in atmospheric CO2. Water Resour. Res., 51, 5450–5463, doi:10.1002/2015WR017031.
  16. Milly, P.C.D. and K.A. Dunne, 2016: Potential evapotranspiration and continental drying. Nat. Clim. Chang., 6, 946–949, doi:10.1038/nclimate3046.
  17. Greve, P., M.L. Roderick, and S.I. Seneviratne, 2017: Simulated changes in aridity from the last glacial maximum to 4xCO2. Environ. Res. Lett., 12, 114021, doi:10.1088/1748-9326/aa89a3.
  18. Roderick, M.L., P. Greve, and G.D. Farquhar, 2015: On the assessment of aridity with changes in atmospheric CO2. Water Resour. Res., 51, 5450–5463, doi:10.1002/2015WR017031.
  19. Greve, P., M.L. Roderick, and S.I. Seneviratne, 2017: Simulated changes in aridity from the last glacial maximum to 4xCO2. Environ. Res. Lett., 12, 114021, doi:10.1088/1748-9326/aa89a3.
  20. Lemordant, L., P. Gentine, A.S. Swann, B.I. Cook, and J. Scheff, 2018: Critical impact of vegetation physiology on the continental hydrologic cycle in response to increasing CO2. Proc. Natl. Acad. Sci. U.S.A., 115, 4093–4098, doi:10.1073/pnas.1720712115.
  21. FAO, 2016: Trees, Forests and Land Use in Drylands: The First Global Assessment Preliminary Findings. Food and Agriculture Organization of the United Nations, Rome, Italy, 31 pp.
  22. Ramankutty, N., A.T. Evan, C. Monfreda, and J.A. Foley, 2008: Farming the planet: 1. Geographic distribution of global agricultural lands in the year 2000. Global Biogeochem. Cycles, 22, GB1003: 1-19, doi:10.1029/2007GB002952.
  23. Fetzel, T. et al., 2017: Quantification of uncertainties in global grazing systems assessment. Global Biogeochem. Cycles, 31, 1089–1102, doi:10.1002/2016GB005601.
  24. Erb, K.-H. et al., 2016: Livestock grazing, the neglected land use. In: Social Ecology [Haberl, H., M. Fischer-Kowalski and F. Winiwarter (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 295–313.
  25. Safriel, U.N., 2007: The assessment of global trends in land degradation. In: Climate and Land Degradation [Sivakumar, M.V.K. and N. Ndiang’ui (eds.)]. Springer, Berlin, Germany, pp. 1–38.
  26. Bai, Z.G., D.L. Dent, L. Olsson, and M.E. Schaepman, 2008: Proxy global assessment of land degradation. Soil Use Manag., 24, 223–234, doi:10.1111/j.1475-2743.2008.00169.x.
  27. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  28. Klein Goldewijk, K., A. Beusen, J. Doelman, and E. Stehfest, 2017: Anthropogenic land use estimates for the Holocene – HYDE 3.2. Earth Syst. Sci. Data, 9, 927–953, doi:10.5194/essd-9-927-2017.
  29. Imeson, A., 2012: Desertification, Land Degradation and Sustainability. Wiley-Blackwell, Oxford, UK. ISBN:9780470714485, DOI:10.1002/9781119
977759, 326 pp.
  30. Sommer, S. et al., 2011: Application of indicator systems for monitoring and assessment of desertification from national to global scales. L. Degrad. Dev, 22, 184–197, doi:10.1002/ldr.1084.
  31. Zucca, C., R. Della Peruta, R. Salvia, S. Sommer, and M. Cherlet, 2012: Towards a World Desertification Atlas. Relating and selecting indicators and data sets to represent complex issues. Ecol. Indic., 15, 157–170, doi:10.1016/J.ECOLIND.2011.09.012.
  32. Bestelmeyer, B.T., M.C. Duniway, D.K. James, L.M. Burkett, and K.M. Havstad, 2013: A test of critical thresholds and their indicators in a desertification-prone ecosystem: More resilience than we thought. Ecol. Lett., 16, 339–345, doi:10.1111/ele.12045.
  33. Maestre, F.T., and A. Escudero, 2009: Is the patch size distribution of vegetation a suitable indicator of desertification processes? Ecology, 90, 1729–1735, doi:10.1890/08-2096.1.
  34. Kéfi, S., C.L. Alados, R.C.G. Chaves, Y. Pueyo, and M. Rietkerk, 2010: Is the patch size distribution of vegetation a suitable indicator of desertification processes? Comment. Ecology, 91, 3739–3742, doi:10.1890/09-1915.1.
  35. Piao, S., J. Fang, H. Liu, and B. Zhu, 2005: NDVI-indicated decline in desertification in China in the past two decades. Geophys. Res. Lett., 32, L06402, doi:10.1029/2004GL021764.
  36. Bestelmeyer, B.T., M.C. Duniway, D.K. James, L.M. Burkett, and K.M. Havstad, 2013: A test of critical thresholds and their indicators in a desertification-prone ecosystem: More resilience than we thought. Ecol. Lett., 16, 339–345, doi:10.1111/ele.12045.
  37. Baskan, O., O. Dengiz, and İ.T. Demirag, 2017: The land productivity dynamics trend as a tool for land degradation assessment in a Dryland ecosystem. Environ. Monit. Assess., 189, 212, doi:10.1007/s10661-017-5909-3.
  38. Zhou, W., C. Gang, F. Zhou, J. Li, X. Dong, and C. Zhao, 2015: Quantitative assessment of the individual contribution of climate and human factors to desertification in Northwest China using net primary productivity as an indicator. Ecol. Indic., 48, 560–569, doi:10.1016/J.ECOLIND.2014.08.043.
  39. Symeonakis, E., N. Karathanasis, S. Koukoulas, and G. Panagopoulos, 2016: Monitoring sensitivity to land degradation and desertification with the environmentally sensitive area index: The case of Lesvos Island. L. Degrad. Dev., 27, 1562–1573, doi:10.1002/ldr.2285.
  40. Dharumarajan, S., T.F.A. Bishop, R. Hegde, and S.K. Singh, 2018: Desertification vulnerability index – An effective approach to assess desertification processes: A case study in Anantapur District, Andhra Pradesh, India. 
L. Degrad. Dev., 29, 150–161, doi:10.1002/ldr.2850.
  41. Salvati, L., M. Zitti, T. Ceccarelli, and L. Perini, 2009: Developing a synthetic index of land vulnerability to drought and desertification. Geogr. Res., 47, 280–291, doi:10.1111/j.1745-5871.2009.00590.x.
  42. Cherlet, M. et al. (eds.), 2018: World Atlas of Desertification. Publication Office of the European Union, Luxembourg, 248 pp.
  43. Myhre, G., D. Shindell, F.-M. Breon, W. Collins, J. Fuglestvedt, J. Huang, 
D. Koch, J.-F. Lamarque, D. Lee, B. Mendoza, T. Nakajima, A. Robock, 
G. Stephens, T. Takemura, and H. Zhang, 2013: Anthropogenic and Natural Radiative Forcing. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, 
G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdon and New York, NY, USA.
  44. Boucher, O., D. Randall, P. Artaxo, C. Bretherton, G. Feingold, P. Forster, 
V.-M. Kerminen, Y. Kondo, H. Liao, U. Lohmann, P. Rasch, S.K. Satheesh, 
S. Sherwood, B. Stevens and X.Y. Zhang, 2013: Clouds and aerosols. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, 
J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdon and New York, NY, USA, pp. 571–657.
  45. Field, C.B., V.R. Barros, K.J. Mach, M.D. Mastrandrea, M. van Aalst, 
W.N. Adger, D.J. Arent, J. Barnett, R. Betts, T.E. Bilir, J. Birkmann, J. Carmin, 
D.D. Chadee, A.J. Challinor, M. Chatterjee, W. Cramer, D.J. Davidson, Y.O. Estrada, J.-P. Gattuso, Y. Hijioka, O. Hoegh-Guldberg, H.Q. Huang, 
G.E. Insarov, R.N. Jones, R.S. Kovats, P. Romero-Lankao, J.N. Larsen, 
I.J. Losada, J.A. Marengo, R.F. McLean, L.O. Mearns, R. Mechler, J.F. Morton, I. Niang, T. Oki, J.M. Olwoch, M. Opondo, E.S. Poloczanska, H.-O. Pörtner, M.H. Redsteer, A. Reisinger, A. Revi, D.N. Schmidt, M.R. Shaw, W. Solecki, D.A. Stone, J.M.R. Stone, K.M. Strzepek, A.G. Suarez, P. Tschakert, 
R. Valentini, S. Vicuña, A. Villamizar, K.E. Vincent, R. Warren, L.L. White, 
T.J. Wilbanks, P.P. Wong, and G.W. Yohe, 2014: Technical Summary. Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., 
V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, 
M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, E.S. Kissel, A.N. Levy, 
S. MacCracken, P.R. Mastrandrea and L.L.White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, USA.
  46. Field, C.B., V.R. Barros, K.J. Mach, M.D. Mastrandrea, M. van Aalst, 
W.N. Adger, D.J. Arent, J. Barnett, R. Betts, T.E. Bilir, J. Birkmann, J. Carmin, 
D.D. Chadee, A.J. Challinor, M. Chatterjee, W. Cramer, D.J. Davidson, Y.O. Estrada, J.-P. Gattuso, Y. Hijioka, O. Hoegh-Guldberg, H.Q. Huang, 
G.E. Insarov, R.N. Jones, R.S. Kovats, P. Romero-Lankao, J.N. Larsen, 
I.J. Losada, J.A. Marengo, R.F. McLean, L.O. Mearns, R. Mechler, J.F. Morton, I. Niang, T. Oki, J.M. Olwoch, M. Opondo, E.S. Poloczanska, H.-O. Pörtner, M.H. Redsteer, A. Reisinger, A. Revi, D.N. Schmidt, M.R. Shaw, W. Solecki, D.A. Stone, J.M.R. Stone, K.M. Strzepek, A.G. Suarez, P. Tschakert, 
R. Valentini, S. Vicuña, A. Villamizar, K.E. Vincent, R. Warren, L.L. White, 
T.J. Wilbanks, P.P. Wong, and G.W. Yohe, 2014: Technical Summary. Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., 
V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, 
M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, E.S. Kissel, A.N. Levy, 
S. MacCracken, P.R. Mastrandrea and L.L.White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, USA.
  47. Fleurbaey, M., S. Kartha, S. Bolwig, Y.L. Chee, Y. Chen, E. Corbera, F. Lecocq, W. Lutz, M.S. Muylaert, R.B. Norgaard, C. Oker-eke, and A.D. Sagar, 2014: Sustainable Development and Equity. Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Edenhofer O., R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, A. Adler, 
I. Baum, S. Brunner, P. Eickemeier, B. Kriemann, J. Savolainen, S. Schlömer, 
C. von Stechow, T. Zwickel and J.C. Minx (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 238–350.
  48. Hoegh-Guldberg, O., D. Jacob, M. Taylor, M. Bindi, S. Brown, I. Camilloni, A. Diedhiou, R. Djalante, K.L. Ebi, F. Engelbrecht, J.Guiot, Y. Hijioka, S. Mehrotra, A. Payne, S.I. Seneviratne, A. Thomas, R. Warren, and G. Zhou, 2018: Impacts of 1.5°C of Global Warming on Natural and Human Systems. In: Global Warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty [Masson-Delmotte, V., P. Zhai, 
H.-O. Pörtner, D. Roberts, J. Skea, P.R. Shukla, A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, J.B.R. Matthews, Y. Chen, X. Zhou, M.I.Gomis, E. Lonnoy, T.Maycock, M.Tignor, and T. Waterfield (eds.)]. In press. pp. 175–311.
  49. IPBES, 2018: Assessment Report on Land Degradation and Restoration [Montanarella, L., R. Scholes., and A. Brainich. (eds.)]. Secretariate of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, Bonn, Germany, 686 pp.
  50. Cherlet, M. et al. (eds.), 2018: World Atlas of Desertification. Publication Office of the European Union, Luxembourg, 248 pp.
  51. Koutroulis, A.G., 2019: Dryland changes under different levels of global warming. Sci. Total Environ., 655, 482–511, doi:10.1016/J.SCITOTENV.2018.11.215.
  52. van der Esch, S. et al., 2017: Exploring Future Changes in Land Use and Land Condition and the Impacts on Food, Water, Climate Change and Biodiversity: Scenarios for the UNCCD Global Land Outlook. Policy Report, PBL Netherlands Environmental Assessment Agency, the Hague, Netherlands, 115 pp.
  53. van der Esch, S. et al., 2017: Exploring Future Changes in Land Use and Land Condition and the Impacts on Food, Water, Climate Change and Biodiversity: Scenarios for the UNCCD Global Land Outlook. Policy Report, PBL Netherlands Environmental Assessment Agency, the Hague, Netherlands, 115 pp.
  54. Bai, Z.G., D.L. Dent, L. Olsson, and M.E. Schaepman, 2008: Proxy global assessment of land degradation. Soil Use Manag., 24, 223–234, doi:10.1111/j.1475-2743.2008.00169.x.
  55. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  56. van der Esch, S. et al., 2017: Exploring Future Changes in Land Use and Land Condition and the Impacts on Food, Water, Climate Change and Biodiversity: Scenarios for the UNCCD Global Land Outlook. Policy Report, PBL Netherlands Environmental Assessment Agency, the Hague, Netherlands, 115 pp.
  57. UN-EMG, 2011: Global Drylands: A UN System-wide Response. United Nations Environment World Conservation Monitoring Centre, Cambridge, UK, 132 pp.
  58. Rosenzweig, C. et al., 2014: Assessing agricultural risks of climate change in the 21st century in a global gridded crop model intercomparison. Proc. Natl. Acad. Sci. U.S.A., 111, 3268–3273, doi:10.1073/pnas.1222463110.
  59. Schlenker, W., and D.B. Lobell, 2010: Robust negative impacts of climate change on African agriculture. Environ. Res. Lett., 5, 14010, doi:10.1088/1748-9326/5/1/014010.
  60. CGIAR-RPDS, 2014: Annual Report 2014. Pathways to Lasting Impact for Rural Dryland Communities in the Developing World. Amman, Jordan. Consultative Group for International Agricultural Research (CGIAR). 68 pages.
  61. Rass, N., 2006: Policies and Strategies to Address the Vulnerability of Pastoralists in Sub-Saharan Africa. PPLPI Working Paper No. 37, Food and Agriculture Organization of the United Nations, Rome, Italy, 22 pp.
  62. Secretariat of the Convention on Biological Diversity, 2010: Pastoralism, Nature Conservation and Development: A Good Practice Guide. Convention on Biological Diversity, Montreal, Canada, 40 pp.
  63. Dong, S., 2016: Overview: Pastoralism in the World. In: Building Resilience of Human-Natural Systems of Pastoralism in the Developing World [Dong, S., K.A.S. Kassam, J. Tourrand and Boone, R.B. (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 1–37.
  64. Carr-Hill, R., 2013: Missing millions and measuring development progress. World Dev., 46, 30–44, doi:10.1016/J.WORLDDEV.2012.12.017.
  65. Varghese, N., and N.P. Singh, 2016: Linkages between land use changes, desertification and human development in the Thar Desert Region of India. Land Use Policy, 51, 18–25, doi:10.1016/J.LANDUSEPOL.2015.11.001.
  66. Krätli, S., and N. Schareika, 2010: Living off uncertainty: The intelligent animal production of Dryland pastoralists. Eur. J. Dev. Res., 22, 605–622, doi:10.1057/ejdr.2010.41.
  67. Snorek, J., F.G. Renaud, and J. Kloos, 2014: Divergent adaptation to climate variability: A case study of pastoral and agricultural societies in Niger. Glob. Environ. Chang., 29, 371–386, doi:10.1016/J.GLOENVCHA.2014.06.014.
  68. Abbass, I., 2014: No retreat no surrender conflict for survival between Fulani pastoralists and farmers in Northern Nigeria. Eur. Sci. J., 8: 331–346, 
doi:10.19044/esj.2012.v8n1p%25p. DOI: http://dx.doi.org/10.19044/esj.2012.
v8n1p%25p,331-346.
  69. Dimelu, M., E. Salifu, and E. Igbokwe, 2016: Resource use conflict in agrarian communities, management and challenges: A case of farmer-herdsmen conflict in Kogi State, Nigeria. J. Rural Stud., 46, 147–154.
  70. Elhadary, Y., 2014: Examining drivers and indicators of the recent changes among pastoral communities of Butana locality, Gedarif State, Sudan. Am. J. Sociol. Res., 4, 88–101.
  71. Morton, J., 2010: Why should governmentality matter for the study of pastoral development? Nomad. People., 14, 6–30, doi:10.3167/np.2010.140102.
  72. Dong, S., 2016: Overview: Pastoralism in the World. In: Building Resilience of Human-Natural Systems of Pastoralism in the Developing World [Dong, S., K.A.S. Kassam, J. Tourrand and Boone, R.B. (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 1–37.
  73. López-i-Gelats, F., E.D.G. Fraser, J.F. Morton, and M.G. Rivera-Ferre, 2016: What drives the vulnerability of pastoralists to global environmental change? A qualitative meta-analysis. Glob. Environ. Chang., 39, 258–274, doi:10.1016/J.GLOENVCHA.2016.05.011.
  74. Gerber, N., E. Nkonya, and J. von Braun, 2014: Land degradation, poverty and marginality. In: Marginality: Addressing the nexus of poverty, exclusion and ecology [Von Braun, J., and F.W. Gatzweiler (eds.)]. Springer Netherlands, Dordrecht, Netherlands, pp. 181–202.
  75. Barbier, E.B., and J.P. Hochard, 2016: Does land degradation increase poverty in developing countries? PLoS One, 11, e0152973, doi:10.1371/journal.pone.0152973.
  76. World Bank, 2018: Groundswell: Preparing for Internal Climate Migration. World Bank Group, Washington DC, USA, 256 pp.
  77. World Bank, 2018: Groundswell: Preparing for Internal Climate Migration. World Bank Group, Washington DC, USA, 256 pp.
  78. Safriel, U., Adeel, Z. et al., 2005: Dryland systems. In: Ecosystems and Human Well-Being: Current State and Trends [Hassan, R., R.J. Scholes, and N. Ash, (eds.)]. Island Press, Washington, DC, USA, pp. 623–662.
  79. Thornton, P.K., P.J. Ericksen, M. Herrero, and A.J. Challinor, 2014: Climate variability and vulnerability to climate change: A review. Glob. Chang. Biol., 20, 3313–3328, doi:10.1111/gcb.12581.
  80. Fraser, E.D.G., A. Dougill, K. Hubacek, C. Quinn, J. Sendzimir, and M. Termansen, 2011: Assessing vulnerability to climate change in dryland livelihood systems: Conceptual challenges and interdisciplinary solutions. Ecol. Soc., 16 (3): 3. DOI: 10.5751/ES-03402-160303.
  81. Thomas, R., 2008: Opportunities to reduce the vulnerability of Dryland farmers in Central and West Asia and North Africa to climate change. Agric. Ecosyst. Environ., 126, 36–45, doi:10.1016/j.agee.2008.01.011.
  82. Bourguignon, F., and S.R. Chakravarty, 2003: The measurement of multidimensional poverty. J. Econ. Inequal., 1, 25–49, doi:10.1023/
A:1023913831342.
  83. Alkire, S., and M.E. Santos, 2010: Acute Multidimensional Poverty: A New Index for Developing Countries. Working Paper: 38, University of Oxford, Poverty and Human Development Initiative, Oxford, UK, 133 pp.
  84. Alkire, S., and M.E. Santos, 2014: Measuring acute poverty in the developing world: Robustness and scope of the multidimensional poverty index. World Dev., 59, 251–274, doi:10.1016/J.WORLDDEV.2014.01.026.
  85. Sietz, D., M.K.B. Lüdeke and C. Walther, 2011: Categorisation of typical vulnerability patterns in global drylands. Glob. Environ. Chang., 21, 431–440, doi:10.1016/J.GLOENVCHA.2010.11.005.
  86. Reynolds, J.F., 2011: Scientific concepts for an integrated analysis of desertification. L. Degrad. Dev., 22, 166–183, doi:10.1002/ldr.1104.
  87. Safriel, U., and Z. Adeel, 2008: Development paths of drylands: Thresholds and sustainability. Sustain. Sci., 3, 117–123, doi:10.1007/s11625-007-0038-5.
  88. Stafford Smith, M., 2016: Desertification: Reflections on the Mirage. In: The End of Desertification? Disputing Environmental Change in the Drylands [Behnke, R. and M. Mortimore (eds.)]. Springer, Berlin, Germany, pp. 539–560.
  89. Salvati, L., 2014: A socioeconomic profile of vulnerable land to desertification in Italy. Sci. Total Environ., 466–467, 287–299. doi.org/10.1016/j.scitotenv.2013.06.091.
  90. Nyantakyi-Frimpong, H., and R. Bezner-Kerr, 2015: The relative importance of climate change in the context of multiple stressors in semi-arid Ghana. Glob. Environ. Chang., 32, 40–56, doi:10.1016/J.GLOENVCHA.2015.03.003.
  91. Sultana, F., 2014: Gendering climate change: Geographical insights. Prof. Geogr., 66, 372–381, doi:10.1080/00330124.2013.821730.
  92. Rahman, M.S., 2013: Climate change, disaster and gender vulnerability: A study on two divisions of Bangladesh. Am. J. Hum. Ecol., 2, 72–82, doi:10.11634/216796221504315.
  93. Nyantakyi-Frimpong, H., and R. Bezner-Kerr, 2015: The relative importance of climate change in the context of multiple stressors in semi-arid Ghana. Glob. Environ. Chang., 32, 40–56, doi:10.1016/J.GLOENVCHA.2015.03.003.
  94. Sultana, F., 2014: Gendering climate change: Geographical insights. Prof. Geogr., 66, 372–381, doi:10.1080/00330124.2013.821730.
  95. Rahman, M.S., 2013: Climate change, disaster and gender vulnerability: A study on two divisions of Bangladesh. Am. J. Hum. Ecol., 2, 72–82, doi:10.11634/216796221504315.
  96. Safriel, U., and Z. Adeel, 2008: Development paths of drylands: Thresholds and sustainability. Sustain. Sci., 3, 117–123, doi:10.1007/s11625-007-0038-5.
  97. Davies, J., 2017: The Land in Drylands: Thriving In Uncertainty through Diversity. Working Paper, Global Land Outlook, Bonn, Germany, 18 pp.
  98. Davies, J. et al., 2016: Improving Governance of Pastoral Lands: Implementing the Voluntary Guidelines on the Responsible Governance of Tenure Of Land, Fisheries and Forests in the Context of National Food Security. Governance of Tenure Technical Guide No. 6, Food and Agriculture Organization of the United Nations, Rome, 152 pp.
  99. Gari, L., 2006: A history of the Hima conservation system. Environ. Hist. Camb., 12, 213–228, doi:10.3197/096734006776680236.
  100. Hussein, I.A.E., 2011: Desertification process in Egypt. In: Coping with Global Environmental Change, Disasters and Security: Threats, Challenges, Vulnerabilities and Risks [Brauch, H.G., U. Oswald Spring, C. Mesjasz, 
J. Grin, P. Kameri-Mbote, B. Chourou, P. Dunay and J. Brikmann (eds.)]. Springer, Berlin, Germany, pp. 863–874.
  101. Fernandez-Gimenez, M.E., 2000: The role of Mongolian nomadic pastoralists’ ecological knowledge in rangeland management. Ecol. Appl., 10, 1318–1326, doi:10.1890/1051-0761(2000)010[1318:TROMNP]2.0.CO;2.
  102. Dominguez, P., 2014: Current situation and future patrimonializing perspectives for the governance of agropastoral resources in the Ait Ikis transhumants of the High Atlas (Morocco). In: Herrera P, Davies J, Baena P (eds.) The Governance of Rangelands. Collective Action for Sustainable Pastoralism. Routledge, Oxon and New York, pp. 126–144.
  103. Fernández-Giménez, M.E., and F. Fillat Estaque, 2012: Pyrenean Pastoralists’ Ecological knowledge: Documentation and application to natural resource management and adaptation. Hum. Ecol., 40, 287–300, doi:10.1007/s10745-012-9463-x.
  104. Hussein, I.A.E., 2011: Desertification process in Egypt. In: Coping with Global Environmental Change, Disasters and Security: Threats, Challenges, Vulnerabilities and Risks [Brauch, H.G., U. Oswald Spring, C. Mesjasz, 
J. Grin, P. Kameri-Mbote, B. Chourou, P. Dunay and J. Brikmann (eds.)]. Springer, Berlin, Germany, pp. 863–874.
  105. Kodirekkala, K.R., 2017: Internal and external factors affecting loss of traditional knowledge: Evidence from a Horticultural Society in South India. J. Anthropol. Res., 73, 22–42, doi:10.1086/690524.
  106. Moreno-Calles, A.I., A. Casas, E. García-Frapolli, and I. Torres-García, 2012: Traditional agroforestry systems of multi-crop ‘milpa’ and ‘chichipera’ cactus forest in the arid Tehuacán Valley, Mexico: Their management and role in people’s subsistence. Agrofor. Syst., 84, 207–226, doi:10.1007/s10457-011-9460-x.
  107. Cowie, A.L. et al., 2018: Land in balance: The scientific conceptual framework for. Environ. Sci. Policy, 79, 25–35, doi:10.1016/J.ENVSCI.2017.10.011.
  108. Liniger, H., R. Mektaschi Studer, P. Moll, and U. Zander, 2017: Making Sense of Research for Sustainable Land Management. Centre for Development and Environment, University of Bern, Switzerland and Helmholtz-Centre for Environmental Research GmbH-UFZ, Leipzig, Germany. ISBN 978-3-944280-99-8, 304 pp.
  109. Safriel, U., and Z. Adeel, 2008: Development paths of drylands: Thresholds and sustainability. Sustain. Sci., 3, 117–123, doi:10.1007/s11625-007-0038-5.
  110. Stafford-Smith, M. et al., 2017: Integration: The Key to Implementing the Sustainable Development Goals. Sustain. Sci., 12, 911–919, doi:10.1007/s11625-016-0383-3.
  111. IPBES, 2018: Assessment Report on Land Degradation and Restoration [Montanarella, L., R. Scholes., and A. Brainich. (eds.)]. Secretariate of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, Bonn, Germany, 686 pp.
  112. Lal, R., 2016: Encyclopedia of Soil Science. CRC Press, Florida, USA, 2653 pp.
  113. Racine, C.K., 2008: Soil in the environment: Crucible of terrestrial life: By Daniel Hillel. Integr. Environ. Assess. Manag., 4, 526. https://doi.org/10.1002/ieam.5630040427.
  114. UNCCD, 2017: Global Land Outlook. First. UNCCD, Bonn, Germany, 336 pp.
  115. Lavauden, L., 1927: Les forêts du Sahara. Berger-Levrault, Paris, France, 26 pp.
  116. Geist, H.H.J., and E.F. Lambin, 2004: Dynamic Causal Patterns of Desertification. Bioscience, 54, 817–829, doi:10.1641/0006-3568(2004)054[0817:DCPOD]2.0.CO;2.
  117. Reynolds, J.F. et al., 2007: Global desertification: Building a science for Dryland development. Science, 316, 847–851, doi:10.1126/science.1131634.
  118. Ginoux, P., J.M. Prospero, T.E. Gill, N.C. Hsu, and M. Zhao, 2012: Global-scale attribution of anthropogenic and natural dust sources and their emission rates based on MODIS Deep Blue aerosol products. Rev. Geophys., 50, doi:10.1029/2012RG000388.
  119. García-Ruiza, J.M., S. Beguería, E. Nadal-Romero, J.C. González-Hidalgo, N. Lana-Renault, and Y. Sanjuán, 2015: A meta-analysis of soil erosion rates across the world. Geomorphology, 239, 160–173, doi:10.1016/j.geomorph.2015.03.008.
  120. Boix-Fayos, C., M. Martínez-Mena, E. Arnau-Rosalén, A. Calvo-Cases, 
V. Castillo, and J. Albaladejo, 2006: Measuring soil erosion by field plots: Understanding the sources of variation. Earth-Science Rev., 78, 267–285, doi:10.1016/J.EARSCIREV.2006.05.005.
  121. FAO, 2015: Global Soil Status, Processes and Trends. Food and Agriculture Organization of the United Nations, Rome, Italy, 605 pp.
  122. Panthou, G., T. Vischel, and T. Lebel, 2014: Recent trends in the regime of extreme rainfall in the Central Sahel. Int. J. Climatol., 34, 3998–4006, doi:10.1002/joc.3984.
  123. Nearing, M.A., C.L. Unkrich, D.C. Goodrich, M.H. Nichols, and T.O. Keefer, 2015: Temporal and elevation trends in rainfall erosivity on a 149 km2 watershed in a semi-arid region of the American Southwest. Int. Soil Water Conserv. Res., 3, 77–85, doi:10.1016/j.iswcr.2015.06.008.
  124. Jish Prakash, P., G. Stenchikov, S. Kalenderski, S. Osipov, and H. Bangalath, 2015: The impact of dust storms on the Arabian Peninsula and the Red Sea. Atmos. Chem. Phys., 15, 199–222, doi:10.5194/acp-15-199-2015.
  125. Klingmüller, K., A. Pozzer, S. Metzger, G.L. Stenchikov, and J. Lelieveld, 2016: Aerosol Optical Depth trend over the Middle East. Atmos. Chem. Phys., 16, 5063–5073, doi:10.5194/acp-16-5063-2016.
  126. Austin, A.T. et al., 2004: Water pulses and biogeochemical cycles in arid and semi-arid ecosystems. Oecologia, 141, 221–235, doi:10.1007/s00442-004-1519-1.
  127. Lal, R., 2009: Sequestering carbon in soils of arid ecosystems. L. Degrad. Dev., 20, 441–454, doi:10.1002/ldr.934.
  128. Lal, R., 2009: Sequestering carbon in soils of arid ecosystems. L. Degrad. Dev., 20, 441–454, doi:10.1002/ldr.934.
  129. Knight, J.R., C.K. Folland, and A.A. Scaife, 2006: Climate impacts of the Atlantic Multidecadal Oscillation. Geophys. Res. Lett., 33, L17706, doi:10.1029/2006GL026242.
  130. Gonzalez-Martin, C., N. Teigell-Perez, B. Valladares, and D.W. Griffin, 2014: The global dispersion of pathogenic microorganisms by dust storms and its relevance to agriculture. Adv. Agron., 127, 1–41, doi:10.1016/B978-0-12-800131-8.00001-7.
  131. Sheen, K.L., D.M. Smith, N.J. Dunstone, R. Eade, D.P. Rowell, and M. Vellinga, 2017: Skilful prediction of Sahel summer rainfall on inter-annual and multi-year time scales. Nat. Commun., 8, 14966, doi:10.1038/ncomms14966.
  132. Pomposi, C., A. Giannini, Y. Kushnir, and D.E. Lee, 2016: Understanding Pacific ocean influence on interannual precipitation variability in the Sahel. Geophys. Res. Lett., 43, 9234–9242, doi:10.1002/2016GL069980.
  133. Hoerling, M. et al., 2006: Detection and attribution of 20th century northern and southern African. Clim. Res., 19, 3989–4008, doi:10.1175/JCLI3842.1.
  134. Huber, S., R. Fensholt, and K. Rasmussen, 2011: Water availability as the driver of vegetation dynamics in the African Sahel from 1982 to 2007. Glob. Planet. Change, 76, 186–195, doi:10.1016/J.GLOPLACHA.2011.01.006.
  135. Tierney, J.E., F.S.R. Pausata, and P.B. DeMenocal, 2017: Rainfall regimes of the Green Sahara. Sci. Adv., 3, e1601503, doi:10.1126/sciadv.1601503.
  136. Collins, J.A. et al., 2017: Rapid termination of the African Humid Period triggered by northern high-latitude cooling. Nat. Commun., 8, 1372, doi:10.1038/s41467-017-01454-y.
  137. Otto-Bliesner, B.L. et al., 2014: Coherent changes of south-eastern equatorial and northern African rainfall during the last deglaciation. Science, 346, 1223–1227, doi:10.1126/science.1259531.
  138. Niedermeyer, E.M., M. Prange, S. Mulitza, G. Mollenhauer, E. Schefuß, and 
M. Schulz, 2009: Extratropical forcing of Sahel aridity during Heinrich stadials. Geophys. Res. Lett., 36, L20707, doi:10.1029/2009GL039687.
  139. Rotstayn, L.D., and U. Lohmann, 2002: Tropical rainfall trends and the indirect aerosol effect. J. Clim., 15, 2103–2116, doi:10.1175/1520-0442(2002)015<2103:TRTATI>2.0.CO;2.
  140. Booth, B.B.B., N.J. Dunstone, P.R. Halloran, T. Andrews, and N. Bellouin, 2012: Aerosols implicated as a prime driver of 20th century North Atlantic climate variability. Nature, 484, 228–232, doi:10.1038/nature10946.
  141. Ackerley, D. et al., 2011: Sensitivity of 20th century Sahel rainfall to sulfate aerosol and CO2 forcing. J. Clim., 24, 4999–5014, doi:10.1175/JCLI-D-11-00019.1.
  142. Funk, C. et al., 2018: Examining the role of unusually warm Indo-Pacific sea-surface temperatures in recent African droughts. Q. J. R. Meteorol. Soc., 144, 360–383, doi:10.1002/qj.3266.
  143. Manjoro, M., V. Kakembo, and K.M. Rowntree, 2012: Trends in soil erosion and woody shrub encroachment in Ngqushwa district, Eastern Cape province, South Africa. Environ. Manage., 49, 570–579, doi:10.1007/s00267-012-9810-0.
  144. Pierson, F.B. et al., 2013: Hydrologic and erosion responses of sagebrush steppe following juniper encroachment, wildfire, and tree cutting. Rangel. Ecol. Manag., 66, 274–289, doi:10.2111/REM-D-12-00104.1.
  145. Eldridge, D.J., L. Wang, and M. Ruiz-Colmenero, 2015: Shrub encroachment alters the spatial patterns of infiltration. Ecohydrology, 8, 83–93, doi:10.1002/eco.1490.
  146. Pierson, F.B. et al., 2013: Hydrologic and erosion responses of sagebrush steppe following juniper encroachment, wildfire, and tree cutting. Rangel. Ecol. Manag., 66, 274–289, doi:10.2111/REM-D-12-00104.1.
  147. Miller, R.F., J.C. Chambers, D.A. Pyke, and F.B. Pierson, 2013: A Review of Fire Effects on Vegetation and Soils in the Great Basin Region: Response and Ecological Site Characteristics. US Department of Agriculture, Forest Service, Rocky Mountain Research Station, Colorado, USA, 126 pp.
  148. Vega, J.A., C. Fernández, and T. Fonturbel, 2005: Throughfall, runoff and soil erosion after prescribed burning in gorse shrubland in Galicia (NW Spain). L. Degrad. Dev., 16, 37–51, doi:10.1002/ldr.643.
  149. Nyman, P., G. Sheridan, and P.N.J. Lane, 2010: Synergistic effects of water repellency and macropore flow on the hydraulic conductivity of a burned forest soil, south-east Australia. Hydrol. Process., 24, 2871–2887, doi:10.1002/hyp.7701.
  150. Holden, S.R., A. Gutierrez, and K.K. Treseder, 2013: Changes in soil fungal communities, extracellular enzyme activities, and litter decomposition across a fire chronosequence in Alaskan boreal forests. Ecosystems, 16, 34–46, doi:10.1007/s10021-012-9594-3.
  151. Pourreza, M., S.M. Hosseini, A.A. Safari Sinegani, M. Matinizadeh, and W.A. Dick, 2014: Soil microbial activity in response to fire severity in Zagros oak (Quercus brantii Lindl.) forests, Iran, after one year. Geoderma, 213, 95–102, doi:10.1016/j.geoderma.2013.07.024.
  152. Weber, C.F., J.S. Lockhart, E. Charaska, K. Aho, and K.A. Lohse, 2014: Bacterial composition of soils in ponderosa pine and mixed conifer forests exposed to different wildfire burn severity. Soil Biol. Biochem., 69, 242–250, doi:10.1016/j.soilbio.2013.11.010.
  153. Liu, B. et al., 2016a: Similar estimates of temperature impacts on global wheat yield by three independent methods. Nat. Clim. Chang., 6, 1130–1136, doi:10.1038/nclimate3115.
  154. Jolly, W.M. et al., 2015: Climate-induced variations in global wildfire danger from 1979 to 2013. Nat. Commun., 6, 1–11, doi:10.1038/ncomms8537.
  155. Williams, A.P. et al., 2010: Forest responses to increasing aridity and warmth in the south-western United States. Proc. Natl. Acad. Sci., 107, 21289–21294, doi:10.1073/pnas.0914211107.
  156. Clarke, H., and J.P. Evans, 2018: Exploring the future change space for fire weather in Southeast Australia. Theor. Appl. Climatol., 136, 513–527, doi:10.1007/s00704-018-2507-4.
  157. Bond, W.J., G.F. Midgley, and F.I. Woodward, 2003: What controls South African vegetation – Climate or fire? South African J. Bot., 69, 79–91.
  158. Bond, W.J., and J.E. Keeley, 2005: Fire as a global ‘herbivore’: The ecology and evolution of flammable ecosystems. Trends Ecol. Evol., 20, 387–394, doi:10.1016/j.tree.2005.04.025.
  159. Balch, J.K., B.A. Bradley, C.M. D’Antonio, and J. Gómez-Dans, 2013: Introduced annual grass increases regional fire activity across the arid western USA (1980–2009). Glob. Chang. Biol., 19, 173–183, doi:10.1111/gcb.12046.
  160. Ziese, M. et al., 2014: The GPCC Drought Index – a new, combined and gridded global drought index. Earth Syst. Sci. Data, 6, 285–295, doi:10.5194/essd-6-285-2014.
  161. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  162. Sietz, D., M.K.B. Lüdeke and C. Walther, 2011: Categorisation of typical vulnerability patterns in global drylands. Glob. Environ. Chang., 21, 431–440, doi:10.1016/J.GLOENVCHA.2010.11.005.
  163. Yan, X., and Y.L. Cai, 2015: Multi-scale anthropogenic driving forces of Karst rocky desertification in south-west China. L. Degrad. Dev., 26, 193–200, doi:10.1002/ldr.2209.
  164. Sterk, G., J. Boardman, and A. Verdoodt, 2016: Desertification: History, causes and options for its control. L. Degrad. Dev., 27, 1783–1787, doi:10.1002/ldr.2525.
  165. Varghese, N., and N.P. Singh, 2016: Linkages between land use changes, desertification and human development in the Thar Desert Region of India. Land Use Policy, 51, 18–25, doi:10.1016/J.LANDUSEPOL.2015.11.001.
  166. Cherlet, M. et al. (eds.), 2018: World Atlas of Desertification. Publication Office of the European Union, Luxembourg, 248 pp.
  167. IPBES, 2018: Assessment Report on Land Degradation and Restoration [Montanarella, L., R. Scholes., and A. Brainich. (eds.)]. Secretariate of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, Bonn, Germany, 686 pp.
  168. UNCCD, 2017: Global Land Outlook. First. UNCCD, Bonn, Germany, 336 pp.
  169. Bestelmeyer, B.T., G.S. Okin, M.C. Duniway, S.R. Archer, N.F. Sayre, 
J.C. Williamson, and J.E. Herrick, 2015: Desertification, land use, and the transformation of global drylands. Front. Ecol. Environ., 13, 28–36, doi:10.1890/140162.
  170. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  171. Lambin, E.F., 2012: Global land availability: Malthus versus Ricardo. Glob. Food Sec., 1, 83–87, doi:10.1016/J.GFS.2012.11.002.
  172. Lambin, E.F. et al., 2013: Estimating the world’s potentially available cropland using a bottom-up approach. Glob. Environ. Chang., 23, 892–901, doi:10.1016/J.GLOENVCHA.2013.05.005.
  173. Eitelberg, D.A., J. van Vliet, and P.H. Verburg, 2015: A review of global potentially available cropland estimates and their consequences for model-based assessments. Glob. Chang. Biol., 21, 1236–1248, doi:10.1111/gcb.12733.
  174. Gutiérrez-Elorza, M., 2006: Erosión e influencia del cambio climático. Rev. C G, 20, 45–59.
  175. Kapović Solomun, M., N. Barger, A. Cerda, S. Keesstra, and M. Marković, 2018: Assessing land condition as a first step to achieving land degradation neutrality: A case study of the Republic of Srpska. Environ. Sci. Policy, 90, 19–27, doi:10.1016/j.envsci.2018.09.014.
  176. Borrelli, P. et al., 2017: An assessment of the global impact of 21st century land use change on soil erosion. Nat. Commun., 8, 2013, doi:10.1038/s41467-017-02142-7.
  177. Lambin, E.F., and P. Meyfroidt, 2011: Global land use change, economic globalization, and the looming land scarcity. Proc. Natl. Acad. Sci. U.S.A., 108, 3465–3472, doi:10.1073/pnas.1100480108.
  178. Hertel, T.W., and D.B. Lobell, 2014: Agricultural adaptation to climate change in rich and poor countries: Current modeling practice and potential for empirical contributions. Energy Econ., 46, 562–575, doi:10.1016/J.ENECO.2014.04.014.
  179. Hallegatte, S., and J. Rozenberg, 2017: Climate change through a poverty lens. Nat. Clim. Chang., 7, 250–256, doi:10.1038/nclimate3253.
  180. Gerber, N., E. Nkonya, and J. von Braun, 2014: Land degradation, poverty and marginality. In: Marginality: Addressing the nexus of poverty, exclusion and ecology [Von Braun, J., and F.W. Gatzweiler (eds.)]. Springer Netherlands, Dordrecht, Netherlands, pp. 181–202.
  181. Way, S.-A., 2016: Examining the links between poverty and land degradation: From blaming the poor toward recognising the rights of the poor. In: Governing Global Desertification: Linking Environmental Degradation, Poverty and Participation [Johnson, P.M. (ed.)]. Routledge, London, UK, pp. 47–62, doi:10.4324/9781315253916-13.
  182. Vu, Q.M., Q.B. Le, E. Frossard, and P.L.G. Vlek, 2014: Socio-economic and biophysical determinants of land degradation in Vietnam: An integrated causal analysis at the national level. Land Use Policy, 36, 605–617, doi:10.1016/J.LANDUSEPOL.2013.10.012.
  183. Chen, R., C. Ye, Y. Cai, X. Xing, and Q. Chen, 2014: The impact of rural out-migration on land use transition in China: Past, present and trend. Land Use Policy, 40, 101–110, doi:10.1016/J.LANDUSEPOL.2013.10.003.
  184. Wang, X., F. Yamauchi, K. Otsuka, and J. Huang, 2014: Wage Growth, Landholding, and Mechanization in Chinese Agriculture. Policy Research Working Paper No. WPS 7138,World Bank Group, Washington, DC, USA, 43 pp.
  185. Wang, X., F. Yamauchi, J. Huang, and S. Rozelle, 2018a: What constrains mechanization in Chinese agriculture? Role of farm size and fragmentation. China Econ. Rev., in press, doi:10.1016/J.CHIECO.2018.09.002.
  186. Taylor, M.J., M. Aguilar-Støen, E. Castellanos, M.J. Moran-Taylor, and K. Gerkin, 2016: International migration, land use change and the environment in Ixcán, Guatemala. Land Use Policy, 54, 290–301, doi:10.1016/J.LANDUSEPOL.2016.02.024.
  187. Gray, C.L., and R.E. Bilsborrow, 2014: Consequences of out-migration for land use in rural Ecuador. Land Use Policy, 36, 182–191, doi:10.1016/J.LANDUSEPOL.2013.07.006.
  188. Qin, H., and T.F. Liao, 2016: Labor out-migration and agricultural change in rural China: A systematic review and meta-analysis. J. Rural Stud., 47, 533–541, doi:10.1016/J.JRURSTUD.2016.06.020.
  189. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  190. Geist, H.H.J., and E.F. Lambin, 2004: Dynamic Causal Patterns of Desertification. Bioscience, 54, 817–829, doi:10.1641/0006-3568(2004)054[0817:DCPOD]2.0.CO;2.
  191. Moussa, B., E. Nkonya, S. Meyer, E. Kato, T. Johnson, and J. Hawkins, 2016: Economics of land degradation and improvement in Niger. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 499–539.
  192. Mythili, G., and J. Goedecke, 2016: Economics of Land Degradation in India. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 431–469.
  193. Sow, S., E. Nkonya, S. Meyer, and E. Kato, 2016: Cost, drivers and action against land degradation in Senegal. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 577–608.
  194. Tun, K.K.K., R.P. Shrestha, and A. Datta, 2015: Assessment of land degradation and its impact on crop production in the dry zone of Myanmar. Int. J. Sustain. Dev. World Ecol., 22, 533–544, doi:10.1080/13504509.2015.1091046.
  195. García-Ruiz, J.M., 2010: The effects of land uses on soil erosion in Spain:
A review. Catena, 81, 1–11, doi:10.1016/j.catena.2010.01.001.
  196. Millennium Ecosystem Assessment, 2005: Ecosystems and Human Well-Being: Desertification Synthesis. World Resource Institute, Washington, DC, USA, 26 pp.
  197. Safriel, U., and Z. Adeel, 2008: Development paths of drylands: Thresholds and sustainability. Sustain. Sci., 3, 117–123, doi:10.1007/s11625-007-0038-5.
  198. Reynolds, J.F. et al., 2007: Global desertification: Building a science for Dryland development. Science, 316, 847–851, doi:10.1126/science.1131634.
  199. Safriel, U., and Z. Adeel, 2008: Development paths of drylands: Thresholds and sustainability. Sustain. Sci., 3, 117–123, doi:10.1007/s11625-007-0038-5.
  200. Reynolds, J.F. et al., 2007: Global desertification: Building a science for Dryland development. Science, 316, 847–851, doi:10.1126/science.1131634.
  201. Behnke, R. and M. Mortimore, 2016: Introduction: The end of desertification? In: The End of Desertification? Disputing Environmental Change in the Drylands [Behnke R. and M. Mortimore (eds.)]. Springer, Berlin, Heidelberg, pp. 1–34.
  202. Swift J., 1996: Desertification: Narratives, winners and losers. In: The Lie of the Land – Challenging Received Wisdom on the African Environment [Leach, M., and R. Mearns (ed.)]. International African Institute, School of Oriental and African Studies, London, UK, pp. 73–90.
  203. Mortimore, M., 2016: Changing paradigms for people-centred development in the Sahel. In: The End of Desertification? Disputing Environmental Change in the Drylands [Behnke, R., and M. Mortimore (eds.)]. Springer, Berlin, Germany, pp. 65–98.
  204. Bestelmeyer, B.T. et al., 2018: The grassland–shrubland regime shift in the south-western United States: Misconceptions and their implications for management. Bioscience, 68, 678–690, doi:10.1093/biosci/biy065.
  205. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  206. Geist, H.H.J., and E.F. Lambin, 2004: Dynamic Causal Patterns of Desertification. Bioscience, 54, 817–829, doi:10.1641/0006-3568(2004)054[0817:DCPOD]2.0.CO;2.
  207. Kok, M. et al., 2016: A new method for analysing socio-ecological patterns of vulnerability. Reg. Environ. Chang., 16, 229–243, doi:10.1007/s10113-014-0746-1.
  208. Polley, H.W. et al., 2013: Climate change and North American rangelands: Trends, projections, and implications. Rangel. Ecol. Manag., 66, 493–511, doi:10.2111/REM-D-12-00068.1.
  209. Ravi, S., D.D. D Breshears, T.E. Huxman, P. D’Odorico,2010: Land degradation in drylands: Interactions among hydrologic-aeolian erosion and vegetation dynamics. Geomorphology, 116, 236–245, doi:10.1016/J.GEOMORPH.2009.11.023.
  210. Scholes, R.J., 2009: Syndromes of dryland degradation in southern Africa. African J. Range Forage Sci., 26, 113–125, doi:10.2989/AJRF.2009.26.3.2.947.
  211. Sietz, D. et al., 2017: Nested archetypes of vulnerability in African drylands: Where lies potential for sustainable agricultural intensification? Environ. Res. Lett., 12, 95006, doi:10.1088/1748-9326/aa768b.
  212. Sietz, D., M.K.B. Lüdeke and C. Walther, 2011: Categorisation of typical vulnerability patterns in global drylands. Glob. Environ. Chang., 21, 431–440, doi:10.1016/J.GLOENVCHA.2010.11.005.
  213. Gibbs, H.K., and J.M. Salmon, 2015: Mapping the world’s degraded lands. Appl. Geogr., 57, 12–21, doi:10.1016/j.apgeog.2014.11.024.
  214. Prince, S.D., 2016: Where Does Desertification Occur? Mapping Dryland Degradation at Regional to Global Scales. Springer, Berlin, Germany, 
pp. 225–263.
  215. Cherlet, M. et al. (eds.), 2018: World Atlas of Desertification. Publication Office of the European Union, Luxembourg, 248 pp.
  216. UNEP, 1992: World Atlas of Desertification [Middleton, N. and D.S.G. Thomas (eds.)]. UNEP.Edward Arnold, London, UK, 69 pp.
  217. Feng, S., and Q. Fu, 2013: Expansion of global drylands under a warming climate. Atmos. Chem. Phys., 13, 10081–10094, doi:10.5194/acp-13-10081-2013.
  218. Asadi Zarch, M.A., B. Sivakumar, and A. Sharma, 2015: Assessment of global aridity change. J. Hydrol., 520, 300–313, doi:10.1016/J.JHYDROL.2014.11.033.
  219. Ji, M., J. Huang, Y. Xie, and J. Liu, 2015: Comparison of Dryland climate change in observations and CMIP5 simulations. Adv. Atmos. Sci., 32, 1565–1574, doi:10.1007/s00376-015-4267-8.
  220. Spinoni, J., J. Vogt, G. Naumann, H. Carrao, and P. Barbosa, 2015: Towards identifying areas at climatological risk of desertification using the Köppen-Geiger classification and FAO Aridity Index. Int. J. Climatol., 35, 2210–2222, doi:10.1002/joc.4124.
  221. Huang, J., H. Yu, X. Guan, G. Wang, and R. Guo, 2016a: Accelerated Dryland expansion under climate change. Nat. Clim. Chang., 6, 166–171, doi:10.1038/nclimate2837.
  222. Ramarao, M.V.S., J. Sanjay, R. Krishnan, M. Mujumdar, A. Bazaz, and A. Revi, 2018: On observed aridity changes over the semi-arid regions of India in a warming climate. Theor. Appl. Climatol., 1–10, doi:10.1007/s00704-018-2513-6.
  223. Roderick, M.L., P. Greve, and G.D. Farquhar, 2015: On the assessment of aridity with changes in atmospheric CO2. Water Resour. Res., 51, 5450–5463, doi:10.1002/2015WR017031.
  224. Milly, P.C.D. and K.A. Dunne, 2016: Potential evapotranspiration and continental drying. Nat. Clim. Chang., 6, 946–949, doi:10.1038/nclimate3046.
  225. Greve, P., M.L. Roderick, and S.I. Seneviratne, 2017: Simulated changes in aridity from the last glacial maximum to 4xCO2. Environ. Res. Lett., 12, 114021, doi:10.1088/1748-9326/aa89a3.
  226. Liu, J. et al., 2017: Water scarcity assessments in the past, present, and future. Earth’s Futur., 5, 545–559, doi:10.1002/2016EF000518.
  227. Sheffield, J., E.F. Wood, and M.L. Roderick, 2012: Little change in global drought over the past 60 years. Nature, 491, 435–438, doi:10.1038/nature11575.
  228. Greve, P., B. Orlowsky, B. Mueller, J. Sheffield, M. Reichstein, and 
S.I. Seneviratne, 2014: Global assessment of trends in wetting and drying over land. Nat. Geosci., 7, 716–721, doi:10.1038/ngeo2247.
  229. Donohue, R.J., M.L. Roderick, T.R. McVicar, and G.D. Farquhar, 2013: Impact of CO2 fertilization on maximum foliage cover across the globe’s warm, arid environments. Geophys. Res. Lett., 40, 3031–3035, doi:10.1002/grl.50563.
  230. Roderick, M.L., P. Greve, and G.D. Farquhar, 2015: On the assessment of aridity with changes in atmospheric CO2. Water Resour. Res., 51, 5450–5463, doi:10.1002/2015WR017031.
  231. Greve, P., M.L. Roderick, and S.I. Seneviratne, 2017: Simulated changes in aridity from the last glacial maximum to 4xCO2. Environ. Res. Lett., 12, 114021, doi:10.1088/1748-9326/aa89a3.
  232. Fensholt, R. et al., 2012: Greenness in semi-arid areas across the globe 1981–2007 – An earth observing satellite based analysis of trends and drivers. Remote Sens. Environ., 121, 144–158, doi:10.1016/J.RSE.2012.01.017.
  233. Andela, N., Y.Y. Liu, A.I.J.M. van Dijk, R.A.M. de Jeu, and T.R. McVicar, 2013: Global changes in Dryland vegetation dynamics (1988-2008) assessed by satellite remote sensing: Comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences, 10, 6657–6676, doi:10.5194/bg-10-6657-2013.
  234. Feng, S., Q. Hu, W. Huang, C.-H. Ho, R. Li, and Z. Tang, 2014: Projected climate regime shift under future global warming from multi-model, multi-scenario CMIP5 simulations. Glob. Planet. Change, 112, 41–52, doi:10.1016/J.GLOPLACHA.2013.11.002.
  235. Spinoni, J., J. Vogt, G. Naumann, H. Carrao, and P. Barbosa, 2015: Towards identifying areas at climatological risk of desertification using the Köppen-Geiger classification and FAO Aridity Index. Int. J. Climatol., 35, 2210–2222, doi:10.1002/joc.4124.
  236. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  237. Gibbs, H.K., and J.M. Salmon, 2015: Mapping the world’s degraded lands. Appl. Geogr., 57, 12–21, doi:10.1016/j.apgeog.2014.11.024.
  238. Vogt, J.V., U. Safriel, G. Von Maltitz, Y. Sokona, R. Zougmore, G. Bastin, and J. Hill, 2011: Monitoring and assessment of land degradation and desertification: Towards new conceptual and integrated approaches. L. Degrad. Dev., 22, 150–165, doi:10.1002/ldr.1075.
  239. Prince, S.D., 2016: Where Does Desertification Occur? Mapping Dryland Degradation at Regional to Global Scales. Springer, Berlin, Germany, 
pp. 225–263.
  240. Cherlet, M. et al. (eds.), 2018: World Atlas of Desertification. Publication Office of the European Union, Luxembourg, 248 pp.
  241. Sonneveld, B.G.J.S., and D.L. Dent, 2007: How good is GLASOD? J. Environ. Manage., 90, 274–283, doi:10.1016/J.JENVMAN.2007.09.008.
  242. Helldén, U., and C. Tottrup, 2008: Regional desertification: A global synthesis. Glob. Planet. Change, 64, 169–176, doi:10.1016/j.gloplacha.2008.10.006.
  243. Hellmann, J.J., J.E. Byers, B.G. Bierwagen, and J.S. Dukes, 2008: Five potential consequences of climate change for invasive species. Conserv. Biol., 22, 534–543, doi:10.1111/j.1523-1739.2008.00951.x.
  244. Sonneveld, B.G.J.S., and D.L. Dent, 2007: How good is GLASOD? J. Environ. Manage., 90, 274–283, doi:10.1016/J.JENVMAN.2007.09.008.
  245. Reed, M.S. et al., 2011: Cross-scale monitoring and assessment of land degradation and sustainable land management: A methodological framework for knowledge management. L. Degrad. Dev., 22, 261–271, doi:10.1002/ldr.1087.
  246. Bai, Z.G., D.L. Dent, L. Olsson, and M.E. Schaepman, 2008: Proxy global assessment of land degradation. Soil Use Manag., 24, 223–234, doi:10.1111/j.1475-2743.2008.00169.x.
  247. de Jong, R., S. de Bruin, A. de Wit, M.E. Schaepman, and D.L. Dent, 2011: Analysis of monotonic greening and browning trends from global NDVI time-series. Remote Sens. Environ., 115, 692–702, doi:10.1016/J.RSE.2010.10.011.
  248. Fensholt, R. et al., 2012: Greenness in semi-arid areas across the globe 1981–2007 – An earth observing satellite based analysis of trends and drivers. Remote Sens. Environ., 121, 144–158, doi:10.1016/J.RSE.2012.01.017.
  249. Andela, N., Y.Y. Liu, A.I.J.M. van Dijk, R.A.M. de Jeu, and T.R. McVicar, 2013: Global changes in Dryland vegetation dynamics (1988-2008) assessed by satellite remote sensing: Comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences, 10, 6657–6676, doi:10.5194/bg-10-6657-2013.
  250. Fensholt, R. et al., 2015: Assessment of vegetation trends in drylands from time series of earth observation data. In: Remote Sensing and Digital Image Processing, Remote Sensing Time Series: Revealing Land Surface Dynamics [Kuenzer, C., S. Dech and W. Wagner (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 159–182.
  251. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  252. Pattison, R.R., J.C. Jorgenson, M.K. Raynolds, and J.M. Welker, 2015: Trends in NDVI and Tundra community composition in the Arctic of NE Alaska between 1984 and 2009. Ecosystems, 18, 707–719, doi:10.1007/s10021-015-9858-9.
  253. Aynekulu, E. et al., 2017: Review of Methodologies for Land Degradation Neutrality Baselines Sub-National Case Studies from Costa Rica and Namibia. CIAT Publication No. 441, International Center for Tropical Agriculture and World Agroforestry Center, Nairobi, Kenya, 58 pp.
  254. Wessels, K.J., F. van den Bergh, and R.J. Scholes, 2012: Limits to detectability of land degradation by trend analysis of vegetation index data. Remote Sens. Environ., 125, 10–22, doi:10.1016/J.RSE.2012.06.022.
  255. de Jong, R., J. Verbesselt, A. Zeileis, and M. Schaepman, 2013: Shifts in global vegetation activity trends. Remote Sens., 5, 1117–1133, doi:10.3390/rs5031117.
  256. Higginbottom, T., and E. Symeonakis, 2014: Assessing land degradation 
and desertification using vegetation index data: Current frameworks and future directions. Remote Sens., 6, 9552–9575, doi:10.3390/rs6109552.
  257. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  258. Evans, J., and R. Geerken, 2004: Discrimination between climate and human-induced Dryland degradation. J. Arid Environ., 57, 535–554, doi:10.1016/S0140-1963(03)00121-6.
  259. de Jong, R., J. Verbesselt, A. Zeileis, and M. Schaepman, 2013: Shifts in global vegetation activity trends. Remote Sens., 5, 1117–1133, doi:10.3390/rs5031117.
  260. Verbesselt, J., H.R., A. Zeileis, and D. Culvenor, 2010: Phenological change detection while accounting for abrupt and gradual trends in satellite image time series. Remote Sens. Environ., 114, 2970–2980, doi:10.1016/J.RSE.2010.08.003.
  261. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  262. Andela, N., Y.Y. Liu, A.I.J.M. van Dijk, R.A.M. de Jeu, and T.R. McVicar, 2013: Global changes in Dryland vegetation dynamics (1988-2008) assessed by satellite remote sensing: Comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences, 10, 6657–6676, doi:10.5194/bg-10-6657-2013.
  263. de Jong, R., J. Verbesselt, A. Zeileis, and M. Schaepman, 2013: Shifts in global vegetation activity trends. Remote Sens., 5, 1117–1133, doi:10.3390/rs5031117.
  264. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  265. Zhang, Y., C. Song, L.E. Band, G. Sun, and J. Li, 2017: Reanalysis of global terrestrial vegetation trends from MODIS products: Browning or greening? Remote Sens. Environ., 191, 145–155, doi:10.1016/J.RSE.2016.12.018.
  266. Wu, W., 2014: The generalized difference vegetation Index (GDVI) for Dryland characterization. Remote Sens., 6, 1211–1233, doi:10.3390/rs6021211.
  267. Ayanu, Y., A. Jentsch, D. Müller-Mahn, S. Rettberg, C. Romankiewicz, and T. Koellner, 2015: Ecosystem engineer unleashed: Prosopis juliflora threatening ecosystem services? Reg. Environ. Chang., 15, 155–167, doi:10.1007/s10113-014-0616-x.
  268. Herrmann, S.M., and G.G. Tappan, 2013: Vegetation impoverishment despite greening: A case study from central Senegal. J. Arid Environ., 90, 55–66, doi:10.1016/J.JARIDENV.2012.10.020.
  269. Brandt, M. et al., 2016a: Assessing woody vegetation trends in Sahelian drylands using MODIS based seasonal metrics. Remote Sens. Environ., 183, 215–225, doi:10.1016/J.RSE.2016.05.027.
  270. Brandt, M. et al., 2016b: Woody plant cover estimation in drylands from Earth Observation based seasonal metrics. Remote Sens. Environ., 172, 28–38, doi:10.1016/J.RSE.2015.10.036.
  271. Evans, J.P., and R. Geerken, 2006: Classifying rangeland vegetation type and coverage using a Fourier component based similarity measure. Remote Sens. Environ., 105, 1–8, doi:10.1016/j.rse.2006.05.017.
  272. Geerken, R.A., 2009: An algorithm to classify and monitor seasonal variations in vegetation phenologies and their inter-annual change. ISPRS J. Photogramm. Remote Sens., 64, 422–431, doi:10.1016/J.ISPRSJPRS.2009.03.001.
  273. Geerken, R., B. Zaitchik, and J.P. Evans, 2005: Classifying rangeland vegetation type and coverage from NDVI time series using Fourier filtered cycle similarity. Int. J. Remote Sens., 26, 5535–5554, doi:10.1080/01431160500300297.
  274. Verbesselt, J., H.R., A. Zeileis, and D. Culvenor, 2010: Phenological change detection while accounting for abrupt and gradual trends in satellite image time series. Remote Sens. Environ., 114, 2970–2980, doi:10.1016/J.RSE.2010.08.003.
  275. Verbesselt, J., R. Hyndman, G. Newnham, and D. Culvenor, 2010: Detecting trend and seasonal changes in satellite image time series. Remote Sens. Environ., 114, 106–115, doi:10.1016/J.RSE.2009.08.014.
  276. Andela, N., Y.Y. Liu, A.I.J.M. van Dijk, R.A.M. de Jeu, and T.R. McVicar, 2013: Global changes in Dryland vegetation dynamics (1988-2008) assessed by satellite remote sensing: Comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences, 10, 6657–6676, doi:10.5194/bg-10-6657-2013.
  277. Tian, F., M. Brandt, Y.Y. Liu, K. Rasmussen, and R. Fensholt, 2017: Mapping gains and losses in woody vegetation across global tropical drylands. Glob. Chang. Biol., 23, 1748–1760, doi:10.1111/gcb.13464.
  278. Liu, Y.Y. et al., 2013: Changing climate and overgrazing are decimating Mongolian steppes. PloS One, 8, e57599. 1-6, doi:10.1371/journal.pone.0057599.
  279. Andela, N., Y.Y. Liu, A.I.J.M. van Dijk, R.A.M. de Jeu, and T.R. McVicar, 2013: Global changes in Dryland vegetation dynamics (1988-2008) assessed by satellite remote sensing: Comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences, 10, 6657–6676, doi:10.5194/bg-10-6657-2013.
  280. Gibbs, H.K., and J.M. Salmon, 2015: Mapping the world’s degraded lands. Appl. Geogr., 57, 12–21, doi:10.1016/j.apgeog.2014.11.024.
  281. Cai, X., X. Zhang, and D. Wang, 2011: Land availability for biofuel production. Environ. Sci. Technol., 45, 334–339, doi:10.1021/es103338e.
  282. Zhu, Z. et al., 2016: Greening of the Earth and its drivers. Nat. Clim. Chang., 6, 791–795, doi:10.1038/nclimate3004.
  283. Borrelli, P. et al., 2017: An assessment of the global impact of 21st century land use change on soil erosion. Nat. Commun., 8, 2013, doi:10.1038/s41467-017-02142-7.
  284. Prăvălie, R., 2016: Drylands extent and environmental issues. A global approach. Earth-Science Rev., 161, 259–278, doi:10.1016/J.EARSCIREV.2016.08.003.
  285. Thiombiano, L., and I. Tourino-Soto, 2007: Status and trends in land degradation in Africa. In: Climate and Land Degradation [Sivakumar, M.V. K and N. Ndiang’ui (eds.)]. Springer, Berlin, Germany, pp. 39–53.
  286. Damberg, L., and A. AghaKouchak, 2014: Global trends and patterns of drought from space. Theor. Appl. Climatol., 117, 441–448, doi:10.1007/s00704-013-1019-5.
  287. Marshall, M., C. Funk, and J. Michaelsen, 2012a: Examining evapotranspiration trends in Africa. Clim. Dyn., 38, 1849–1865, doi:10.1007/s00382-012-1299-y.
  288. Oroda, A.S., 2001: Application of remote sensing to early warning for food security and environmental monitoring in the Horn of Africa. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci., XXXIV, 66–72.
  289. Pricope, N.G., G. Husak, D. Lopez-Carr, C. Funk, and J. Michaelsen, 2013: The climate-population nexus in the East African Horn: Emerging degradation trends in rangeland and pastoral livelihood zones. Glob. Environ. Chang., 23, 1525–1541, doi:10.1016/J.GLOENVCHA.2013.10.002.
  290. Gichenje, H., and S. Godinho, 2018: Establishing a land degradation neutrality national baseline through trend analysis of GIMMS NDVI Time-series. 
L. Degrad. Dev., 29, 2985–2997, doi:10.1002/ldr.3067.
  291. Kihiu, E., 2016a: Pastoral Practices, Economics, and Institutions of Sustainable Rangeland Management in Kenya. PhD Thesis, University of Bonn, Bonn, Germany, 167 pp.
  292. Kihiu, E.N., 2016b: Basic capability effect: Collective management of pastoral resources in south-western Kenya. Ecol. Econ., 123, 23–34, doi:10.1016/J.ECOLECON.2016.01.003.
  293. Otuoma, J., J. Kinyamario, W. Ekaya, M. Kshatriya, and M. Nyabenge, 2009: Effects of human-livestock-wildlife interactions on habitat in an Eastern Kenya rangeland. Afr. J. Ecol., 47, 567–573, doi:10.1111/j.1365-2028.2008.01009.x.
  294. Anyamba, A., and C.J. Tucker, 2005: Analysis of Sahelian vegetation dynamics using NOAA-AVHRR NDVI data from 1981–2003. J. Arid Environ., 63, 596–614, doi:10.1016/J.JARIDENV.2005.03.007.
  295. Huber, S., R. Fensholt, and K. Rasmussen, 2011: Water availability as the driver of vegetation dynamics in the African Sahel from 1982 to 2007. Glob. Planet. Change, 76, 186–195, doi:10.1016/J.GLOPLACHA.2011.01.006.
  296. Brandt, M. et al., 2015: Ground- and satellite-based evidence of the biophysical mechanisms behind the greening Sahel. Glob. Chang. Biol., 21, 1610–1620, doi:10.1111/gcb.12807.
  297. Rishmawi, K., S. Prince, K. Rishmawi, and S.D. Prince, 2016: Environmental and anthropogenic degradation of v egetation in the Sahel from 1982 to 2006. Remote Sens., 8, 948, doi:10.3390/rs8110948.
  298. Tian, F. et al., 2016: Remote sensing of vegetation dynamics in drylands: Evaluating vegetation optical depth (VOD) using AVHRR NDVI and in situ green biomass data over West African Sahel. Remote Sens. Environ., 177, 265–276, doi:10.1016/J.RSE.2016.02.056.
  299. Leroux, L., A. Bégué, D. Lo Seen, A. Jolivot, and F. Kayitakire, 2017: Driving forces of recent vegetation changes in the Sahel: Lessons learned from regional and local level analyses. Remote Sens. Environ., 191, 38–54, doi:10.1016/J.RSE.2017.01.014.
  300. Herrmann, S.M., and C.F. Hutchinson, 2005: The changing contexts of the desertification debate. J. Arid Environ., 63, 538–555, doi:10.1016/J.JARIDENV.2005.03.003.
  301. Damberg, L., and A. AghaKouchak, 2014: Global trends and patterns of drought from space. Theor. Appl. Climatol., 117, 441–448, doi:10.1007/s00704-013-1019-5.
  302. Traore, S.B. et al., 2014: AGRHYMET: A drought monitoring and capacity building center in the West Africa region. Weather Clim. Extrem., 3, 22–30, doi:10.1016/J.WACE.2014.03.008.
  303. Thomas, N., and S. Nigam, 2018: 20th century climate change over Africa: Seasonal hydroclimate trends and Sahara desert expansion. J. Clim., 31, 3349–3370, doi:10.1175/JCLI-D-17-0187.1.
  304. Dimobe, K., A. Ouédraogo, S. Soma, D. Goetze, S. Porembski, and 
A. Thiombiano, 2015: Identification of driving factors of land degradation and deforestation in the wildlife reserve of Bontioli (Burkina Faso, West Africa). Glob. Ecol. Conserv., 4, 559–571, doi:10.1016/j.gecco.2015.10.006.
  305. Hiernaux, P. et al., 2009: Trends in productivity of crops, fallow and rangelands in Southwest Niger: Impact of land use, management and variable rainfall. J. Hydrol., 375, 65–77, doi:10.1016/J.JHYDROL.2009.01.032.
  306. Aladejana, O.O., A.T. Salami, and O.-I.O. Adetoro, 2018: Hydrological responses to land degradation in the Northwest Benin Owena River Basin, Nigeria. J. Environ. Manage., 225, 300–312, doi:10.1016/J.JENVMAN.2018.07.095.
  307. Le, Q.B., L. Tamene and P.L.G. Vlek, 2012: Multi-pronged assessment of land degradation in West Africa to assess the importance of atmospheric fertilization in masking the processes involved. Glob. Planet. Change, 92–93, 71–81, doi:10.1016/J.GLOPLACHA.2012.05.003.
  308. Helldén, U., and C. Tottrup, 2008: Regional desertification: A global synthesis. Glob. Planet. Change, 64, 169–176, doi:10.1016/j.gloplacha.2008.10.006.
  309. Fensholt, R. et al., 2012: Greenness in semi-arid areas across the globe 1981–2007 – An earth observing satellite based analysis of trends and drivers. Remote Sens. Environ., 121, 144–158, doi:10.1016/J.RSE.2012.01.017.
  310. Smith, P. et al., 2013: How much land-based greenhouse gas mitigation can be achieved without compromising food security and environmental goals? Glob. Chang. Biol., 19, 2285–2302, doi:10.1111/gcb.12160.
  311. Herrmann, S.M., and G.G. Tappan, 2013: Vegetation impoverishment despite greening: A case study from central Senegal. J. Arid Environ., 90, 55–66, doi:10.1016/J.JARIDENV.2012.10.020.
  312. Kaptué, A.T., L. Prihodko, and N.P. Hanan, 2015: On regreening and degradation in Sahelian watersheds. Proc. Natl. Acad. Sci. U.S.A., 112, 12133–12138, doi:10.1073/pnas.1509645112.
  313. Herrmann, S.M., and T.K. Sop, 2016: The map is not the territory: How satellite remote sensing and ground evidence have re-shaped the image of Sahelian desertification. In: The End of Desertification? [Behnke, R. and M. Mortimore (eds.)]. Springer Earth System Sciences, Springer, Berlin, Germany, pp. 117–145.
  314. Saha, M.V., T.M. Scanlon, and P. D’Odorico, 2015: Examining the linkage between shrub encroachment and recent greening in water-limited southern Africa. Ecosphere, 6, art156, doi:10.1890/ES15-00098.1.
  315. Weinzierl, T., J. Wehberg, J. Böhner, and O. Conrad, 2016: Spatial Assessment of Land Degradation Risk for the Okavango River Catchment, southern Africa. L. Degrad. Dev., 27, 281–294, doi:10.1002/ldr.2426.
  316. Hirche, A., M. Salamani, A. Abdellaoui, S. Benhouhou, and J.M. Valderrama, 2011: Landscape changes of desertification in arid areas: The case of south-west Algeria. Environ. Monit. Assess., 179, 403–420, doi:10.1007/s10661-010-1744-5.
  317. Ayoub, A.T., 1998: Extent, severity and causative factors of land degradation in the Sudan. J. Arid Environ., 38, 397–409, doi:10.1006/JARE.1997.0346.
  318. Dawelbait, M., and F. Morari, 2012: Monitoring desertification in a savanna region in Sudan using Landsat images and spectral mixture analysis. J. Arid Environ., 80, 45–55, doi:10.1016/j.jaridenv.2011.12.011.
  319. del Barrio, G. et al., 2016: Land degradation states and trends in the north-western Maghreb drylands, 1998–2008. Remote Sens., 8, 603, doi:10.3390/rs8070603.
  320. Kouba, Y., M. Gartzia, A. El Aich, and C.L. Alados, 2018: Deserts do not advance, they are created: Land degradation and desertification in semi-arid environments in the Middle Atlas, Morocco. J. Arid Environ., 158, 1–8, doi:10.1016/J.JARIDENV.2018.07.002.
  321. Lahlaoi, H. et al., 2017: Desertification assessment using MEDALUS model in watershed Oued El Maleh, Morocco. Geosciences, 7, 50, doi:10.3390/geosciences7030050.
  322. Réquier-Desjardins, M., and M. Bied-Charreton, 2006: Évaluation économique des coûts économiques et sociaux de la désertification en Afrique. Centre d’Economie et d’Ethique pour l’Environnement et le Développement, Université de Versailles St Quentin- en- Yvelines, Paris, France, 163 pp.
  323. Prăvălie, R., 2016: Drylands extent and environmental issues. A global approach. Earth-Science Rev., 161, 259–278, doi:10.1016/J.EARSCIREV.2016.08.003.
  324. Miao, L. et al., 2015a: Footprint of research in desertification management in China. L. Degrad. Dev., 26, 450–457, doi:10.1002/ldr.2399.
  325. Lal, R., Stewart, B.A. (eds.), 2012: Advances in Soil Science: Soil Water and Agronomic Productivity. CRC Press, Florida, USA, 320 pp.
  326. Aslam, M., and S.A. Prathapar, 2006: Strategies to Mitigate Secondary Salinization in the Indus Basin of Pakistan: A Selective Review. Research Report 97, International Water Management Institute, Colombo, Sri Lanka, 31 pp.
  327. Chengrui, M., and H.E. Dregne, 2001: Review article: Silt and the future development of China’s Yellow River. Geogr. J., 167, 7–22, doi:10.1111/1475-4959.00002.
  328. Zhou, D., Z. Lin, L. Liu, and D. Zimmermann, 2013: Assessing secondary soil salinization risk based on the PSR sustainability framework. J. Environ. Manage., 128, 642–654, doi:10.1016/J.JENVMAN.2013.06.025.
  329. Cai, X., D.C. McKinney, and M.W. Rosegrant, 2003: Sustainability analysis for irrigation water management in the Aral Sea region. Agric. Syst., 76, 1043–1066, doi:10.1016/S0308-521X(02)00028-8.
  330. Pankova, E.I., 2016: Salinization of irrigated soils in the Middle-Asian Region: Old and new issues. Arid Ecosyst., 6, 241–248, doi:10.1134/S2079096116040077.
  331. Qadir, M., A.D. Noble, A.S. Qureshi, R.K. Gupta, T. Yuldashev, and A. Karimov, 2009: Salt-induced land and water degradation in the Aral Sea Basin: A challenge to sustainable agriculture in Central Asia. Nat. Resour. Forum, 33, 134–149, doi:10.1111/j.1477-8947.2009.01217.x.
  332. Helldén, U., and C. Tottrup, 2008: Regional desertification: A global synthesis. Glob. Planet. Change, 64, 169–176, doi:10.1016/j.gloplacha.2008.10.006.
  333. Chen, D. et al., 2015a: Patterns and drivers of soil microbial communities along a precipitation gradient on the Mongolian Plateau. Landsc. Ecol., 30, 1669–1682, doi:10.1007/s10980-014-9996-z.
  334. Wang, H., Y. Pan, Y. Chen, and Z. Ye, 2017a: Linear trend and abrupt changes of climate indices in the arid region of north-western China. Atmos. Res., 196, 108–118. doi.org/10.1016/j.atmosres.2017.06.008.
  335. Yang, F., and C. Lu, 2015: Spatiotemporal variation and trends in rainfall erosivity in China’s Dryland region during 1961–2012. Catena, 133, 362–372, doi:10.1016/j.catena.2015.06.005.
  336. Zhang, J. et al., 2015: Effects of sand dune stabilization on the spatial pattern of Artemisia ordosica population in Mu Us desert, Northwest China. PLoS One, 10, e0129728, doi:10.1371/journal.pone.0129728.
  337. Xue, Z., Z. Qin, F. Cheng, G. Ding, and H. Li, 2017: Quantitative assessment of aeolian desertification dynamics – A case study in North Shanxi of China (1975 to 2015). Sci. Rep., 7, 10460, doi:10.1038/s41598-017-11073-8.
  338. Li, A., J. Wu, and J. Huang, 2012: Distinguishing between human-induced and climate-driven vegetation changes: A critical application of RESTREND in inner Mongolia. Landsc. Ecol., 27, 969–982, doi:10.1007/s10980-012-9751-2.
  339. Fu, Q., B. Li, Y. Hou, X. Bi, and X. Zhang, 2017: Effects of land use and climate change on ecosystem services in Central Asia’s arid regions: A case study in Altay Prefecture, China. Sci. Total Environ., 607–608, 633–646, doi:10.1016/J.SCITOTENV.2017.06.241.
  340. Muyibul, Z., X. Jianxin, P. Muhtar, S. Qingdong, and Z. Run, 2018: Spatiotemporal changes of land use/cover from 1995 to 2015 in an oasis in the middle reaches of the Keriya River, southern Tarim Basin, Northwest China. CATENA, 171, 416–425, doi:10.1016/J.CATENA.2018.07.038.
  341. Xie, Y., J. Gong, P. Sun, and X. Gou, 2014: Oasis dynamics change and its influence on landscape pattern on Jinta oasis in arid China from 1963a to 2010a: Integration of multi-source satellite images. Int. J. Appl. Earth Obs. Geoinf., 33, 181–191, doi:10.1016/j.jag.2014.05.008.
  342. Lamchin, M. et al., 2016: Assessment of land cover change and desertification using remote sensing technology in a local region of Mongolia. Adv. Sp. Res., 57, 64–77, doi:10.1016/J.ASR.2015.10.006.
  343. Liu, Y.Y. et al., 2013: Changing climate and overgrazing are decimating Mongolian steppes. PloS One, 8, e57599. 1-6, doi:10.1371/journal.pone.0057599.
  344. Issanova, G., and J. Abuduwaili, 2017: Relationship between storms and land degradation. In: Aeolian Process as Dust Storms in the Deserts of Central Asia and Kazakhstan [Issanova, G. and J. Abuduwaili (eds.)]. Springer, Singapore, pp. 71–86.
  345. Lioubimtseva, E., 2015: A multi-scale assessment of human vulnerability to climate change in the Aral Sea Basin. Environ. Earth Sci., 73, 719–729, doi:10.1007/s12665-014-3104-1.
  346. Micklin, P., 2016: The future Aral Sea: Hope and despair. Environ. Earth Sci., 75, 844, doi:10.1007/s12665-016-5614-5.
  347. Xi, X., and I.N. Sokolik, 2015: Seasonal dynamics of threshold friction velocity and dust emission in Central Asia. J. Geophys. Res. Atmos., 120, 1536–1564, doi:10.1002/2014JD022471.
  348. Qadir, M., A.D. Noble, A.S. Qureshi, R.K. Gupta, T. Yuldashev, and A. Karimov, 2009: Salt-induced land and water degradation in the Aral Sea Basin: A challenge to sustainable agriculture in Central Asia. Nat. Resour. Forum, 33, 134–149, doi:10.1111/j.1477-8947.2009.01217.x.
  349. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  350. Chen, T. et al., 2019: Disentangling the relative impacts of climate change and human activities on arid and semi-arid grasslands in Central Asia during 1982–2015. Sci. Total Environ., 653, 1311–1325, doi:10.1016/j.scitotenv.2018.11.058.
  351. Mirzabaev, A., M. Ahmed, J. Werner, J. Pender, and M. Louhaichi, 2016a: Rangelands of Central Asia: Challenges and opportunities. J. Arid Land, 8, 93–108, doi:10.1007/s40333-015-0057-5.
  352. Damberg, L., and A. AghaKouchak, 2014: Global trends and patterns of drought from space. Theor. Appl. Climatol., 117, 441–448, doi:10.1007/s00704-013-1019-5.
  353. UNEP-GEF, 2008: Desertification, Rangelands and Water Resources Working Group Final Thematic Report, 47 pp.
  354. Irshad, M. et al., 2007: Land desertification – An emerging threat to environment and food security of Pakistan. J. Appl. Sci., 7, 1199–1205, doi:10.3923/jas.2007.1199.1205.
  355. Lal, R., 2018: Managing agricultural soils of Pakistan for food and climate. Soil Environ., 37, 1–10.
  356. Kundu, A., N.R. Patel, S.K. Saha, and D. Dutta, 2017: Desertification in western Rajasthan (India): An assessment using remote sensing derived rain-use efficiency and residual trend methods. Nat. Hazards, 86, 297–313, doi:10.1007/s11069-016-2689-y.
  357. Dharumarajan, S., T.F.A. Bishop, R. Hegde, and S.K. Singh, 2018: Desertification vulnerability index – An effective approach to assess desertification processes: A case study in Anantapur District, Andhra Pradesh, India. 
L. Degrad. Dev., 29, 150–161, doi:10.1002/ldr.2850.
  358. Christian, B.A., P.S. Dhinwa, and Ajai, 2018: Long term monitoring and assessment of desertification processes using medium and high resolution satellite data. Appl. Geogr., 97, 10–24, doi:10.1016/J.APGEOG.2018.04.010.
  359. Ajai, R.R., A.S. Arya, A.S. Dhinwa, S.K. Pathan, and K.G. Raj, 2009: Desertification/land degradation status mapping of India. Curr. Sci., 97, 1478–1483.
  360. Singh, G., 2009: Salinity-related desertification and management strategies: Indian experience. L. Degrad. Dev., 20, 367–385, doi:10.1002/ldr.933.
  361. Ministry of Energy Industry and Mineral Resources, 2016: Third National Communication of the Kingdom of Saudi Arabia. Submitted to United Nations Framework Convention on Climate Change, Riyadh, Saudi Arabia, pp. 173–174.
  362. Yahiya, A.B., 2012: Environmental degradation and its impact on tourism in Jazan, KSA using remote sensing and GIS. International Journal Environmental Sciences, 3, 421–432, doi:10.6088/ijes.2012030131041.
  363. Hasanean, H., M. Almazroui, H. Hasanean, and M. Almazroui, 2015: Rainfall: Features and variations over Saudi Arabia, a review. Climate, 3, 578–626, doi:10.3390/cli3030578.
  364. Al-Bakri, J.T. et al., 2016: Modelling desertification risk in the northwest of Jordan using geospatial and remote sensing techniques. Geomatics, Nat. Hazards Risk, 7, 531–549, doi:10.1080/19475705.2014.945102.
  365. Evans, J., and R. Geerken, 2004: Discrimination between climate and human-induced Dryland degradation. J. Arid Environ., 57, 535–554, doi:10.1016/S0140-1963(03)00121-6.
  366. Geerken, R., and M. Ilaiwi, 2004: Assessment of rangeland degradation and development of a strategy for rehabilitation. Remote Sens. Environ., 90, 490–504, doi:10.1016/J.RSE.2004.01.015.
  367. Moridnejad, A., N. Karimi, and P.A. Ariya, 2015: Newly desertified regions in Iraq and its surrounding areas: Significant novel sources of global dust particles. J. Arid Environ., 116, 1–10, doi:10.1016/J.JARIDENV.2015.01.008.
  368. Amiraslani, F., and D. Dragovich, 2011: Combating desertification in Iran over the last 50 years: An overview of changing approaches. J. Environ. Manage., 92, 1–13, doi:10.1016/j.jenvman.2010.08.012.
  369. Damberg, L., and A. AghaKouchak, 2014: Global trends and patterns of drought from space. Theor. Appl. Climatol., 117, 441–448, doi:10.1007/s00704-013-1019-5.
  370. Donohue, R.J., T.R. McVicar, and M.L. Roderick, 2009: Climate-related trends in Australian vegetation cover as inferred from satellite observations, 1981–2006. Glob. Chang. Biol., 15, 1025–1039, doi:10.1111/j.1365-2486.2008.01746.x.
  371. Burrell, A.L., J.P. Evans, and Y. Liu, 2017: Detecting dryland degradation using Time Series Segmentation and Residual Trend analysis (TSS-RESTREND). Remote Sens. Environ., 197, doi:10.1016/j.rse.2017.05.018.
  372. Rengasamy, P., 2006: World salinization with emphasis on Australia. J. Exp. Bot., 57, 1017–1023, doi:10.1093/jxb/erj108.
  373. Eldridge, D.J., and S. Soliveres, 2014: Are shrubs really a sign of declining ecosystem function? Disentangling the myths and truths of woody encroachment in Australia. Aust. J. Bot., 62, 594–608, doi:10.1071/BT14137.
  374. Zdruli, P., 2011: Desertification in the Mediterranean Region. European Institute of the Mediterranean, Girona, Barcelona, 250–255 pp.
  375. European Court of Auditors (ECA), 2018: Desertfication in Europe. Background paper, European Court of Auditors, Luxembourg, 15 pp.
  376. Prăvălie, R., C. Patriche, and G. Bandoc, 2017: Quantification of land degradation sensitivity areas in southern and central south-eastern Europe. New results based on improving DISMED methodology with new climate data. Catena, 158, 309–320, doi:10.1016/j.catena.2017.07.006.
  377. European Court of Auditors (ECA), 2018: Desertfication in Europe. Background paper, European Court of Auditors, Luxembourg, 15 pp.
  378. Safriel, U., 2009: Deserts and desertification: Challenges but also opportunities. L. Degrad. Dev., 20, 353–366, doi:10.1002/ldr.935.
  379. European Court of Auditors (ECA), 2018: Desertfication in Europe. Background paper, European Court of Auditors, Luxembourg, 15 pp.
  380. López-Bermúdez, F., 1990: Soil erosion by water on the desertification of a semi-arid Mediterranean fluvial basin: The Segura Basin, Spain. Agric. Ecosyst. Environ., 33, 129–145, doi:10.1016/0167-8809(90)90238-9.
  381. Lázaro, R., Y. Cantón, A. Solé-Benet, J. Bevan, R. Alexander, L.G. Sancho, and J. Puigdefábregas, 2008: The influence of competition between lichen colonization and erosion on the evolution of soil surfaces in the Tabernas badlands (SE Spain) and its landscape effects. Geomorphology, 102, 252–266, doi:10.1016/J.GEOMORPH.2008.05.005.
  382. Guerra, C.A., J. Maes, I. Geijzendorffer, and M.J. Metzger, 2016: An assessment of soil erosion prevention by vegetation in Mediterranean Europe: Current trends of ecosystem service provision. Ecol. Indic. 60, 213–222, doi:10.1016/j.ecolind.2015.06.043.
  383. Helldén, U., and C. Tottrup, 2008: Regional desertification: A global synthesis. Glob. Planet. Change, 64, 169–176, doi:10.1016/j.gloplacha.2008.10.006.
  384. Fensholt, R. et al., 2012: Greenness in semi-arid areas across the globe 1981–2007 – An earth observing satellite based analysis of trends and drivers. Remote Sens. Environ., 121, 144–158, doi:10.1016/J.RSE.2012.01.017.
  385. Gunin, P., and E. Pankova, 2004: Contemporary processes of degradation and desertification of ecosystems of the East Asian sector of steppes and forest-steppes. In: Modern Global Changes of the Natural Environment. Scientific World, Moscow, Russia, pp. 389–412.
  386. Kust, G.S., O.V. Andreeva, and D.V. Dobrynin, 2011: Desertification assessment and mapping in the Russian Federation. Arid Ecosyst., 1, 14–28, doi:10.1134/S2079096111010057.
  387. Türkeş. M., 1999: Vulnerability of Turkey to desertification with respect to precipitation and aridity conditions. Turkish J. Eng. Environ. Sci., 23, 363–380.
  388. Türkeş, M., 2003: Spatial and temporal variations in precipitation and aridity index series of Turkey. In: Mediterranean Climate [Bolle, H.J. (ed.)]. Springer, Berlin, Germany, pp. 181–213.
  389. Türkeş, M., 2013: İklim Verileri Kullanılarak Türkiye’nin Çölleşme Haritası Dokümanı Hazırlanması Raporu. Orman ve Su İşleri Bakanlığı, Çölleşme ve Erozyonla Mücadele Genel Müdürlüğü Yayını, Ankara, Turkey, 57 pp.
  390. General Directorate of Combating Desertification and Erosion (ÇEMGM), 2017: Combating Desertification and Erosion Activities in Turkey. General Directorate of Combating Desertification and Erosion, Ministry of Forestry and Water Affairs, Republic of Turkey. Ankara, Turkey, 44 pp.
  391. Pontifes, P.A., P.M. García-Meneses, L. Gómez-Aíza, A.I. Monterroso-Rivas, and M. Caso Chávez, 2018: Land use/land cover change and extreme climatic events in the arid and semi-arid ecoregions of Mexico. ATMÓSFERA, 31, 355–372, doi:10.20937/ATM.2018.31.04.04.
  392. Méndez, M., and V. Magaña, 2010: Regional aspects of prolonged meteorological droughts over Mexico and Central America. J. Clim., 23, 1175–1188, doi:10.1175/2009JCLI3080.1.
  393. Stahle, D.W. et al., 2009: Cool- and warm-season precipitation reconstructions over western New Mexico. J. Clim., 22, 3729–3750, doi:10.1175/2008JCLI2752.1.
  394. Becerril-Pina Rocio et al., 2015: Assessing desertification risk in the semi-arid highlands of central Mexico. J. Arid Environ., 120, 4–13, doi:10.1016/j.jaridenv.2015.04.006.
  395. Noojipady, P., S.D. Prince, and K. Rishmawi, 2015: Reductions in productivity due to land degradation in the drylands of the south-western united states. Ecosyst. Heal. Sustain., 1, 1–15, doi:10.1890/EHS15-0020.1.
  396. Yanoff, S., and E. Muldavin, 2008: Grassland-shrubland transformation and grazing: A century-scale view of a northern Chihuahuan Desert grassland. J. Arid Environ., 72, 1594–1605, doi:10.1016/J.JARIDENV.2008.03.012.
  397. Browning, D.M., and S.R. Archer, 2011: Protection from livestock fails to deter shrub proliferation in a desert landscape with a history of heavy grazing. Ecol. Appl., 21, 1629–1642, doi:10.1890/10-0542.1.
  398. Van Auken, O.W., 2009: Causes and consequences of woody plant encroachment into western North American grasslands. J. Environ. Manage., 90, 2931–2942, doi:10.1016/J.JENVMAN.2009.04.023.
  399. Rachal, D.M., H.C. Monger, G.S. Okin, and D.C. Peters, 2012: Landform influences on the resistance of grasslands to shrub encroachment, Northern Chihuahuan Desert, USA.J. Maps, 8, 507–513, doi:10.1080/17445647.2012.727593.
  400. Potter, C., and J. Weigand, 2016: Analysis of desert sand dune migration patterns from Landsat image time deries for the southern California desert. J. Remote Sens. GIS, 5, 1–8, doi:10.4172/2469-4134.1000164.
  401. Lam, D.K. et al., 2011: Tracking desertification in California using remote sensing: A sand dune encroachment approach. Remote Sens., 3, 1–13, doi:10.3390/rs3010001.
  402. Kertis, C.A., 2003: Soil Erosion on Cropland in The United States: Status and Trends for 1982–2003. Proceedings of the Eight Federal Interagency Sedimentation Conference, April 2006, Nevada, USA.
  403. Morales, C. et al., 2011: Measuring the Economic Value of Land Degradation/Desertification Considering the Effects of Climate Change. A Study for Latin America and the Caribbean. Communication au Séminaire, Politiques, programmes et projets de lutte contre la désertification, quelles évaluations, Montpellier, France. 20 pages.
  404. Verón, S.R., L.J. Blanco, M.A. Texeira, J.G.N. Irisarri, and J.M. Paruelo, 2017: Desertification and ecosystem services supply: The case of the Arid Chaco of South America. J. Arid Environ., 159, 66–74, doi:10.1016/J.JARIDENV.2017.11.001.
  405. Fernández, O.A., M.E. Gil, and R.A. Distel, 2009: The challenge of rangeland degradation in a temperate semi-arid region of Argentina: The Caldenal. 
L. Degrad. Dev., 20, 431–440, doi:10.1002/ldr.851.
  406. Bisigato, A.J., and R.M.L. Laphitz, 2009: Ecohydrological effects of grazing-induced degradation in the Patagonian Monte, Argentina. Austral Ecol., 34, 545–557, doi:10.1111/j.1442-9993.2009.01958.x.
  407. Vieira, R.M.S.P. et al., 2015: Identifying areas susceptible to desertification in the Brazilian Northeast. Solid Earth, 6, 347–360, doi:10.5194/se-6-347-2015.
  408. Leal, I.R., J.M.C. Da Silva, M. Tabarelli, and T.E. Lacher, 2005: Changing the course of biodiversity conservation in the Caatinga of north-eastern Brazil. Conserv. Biol., 19, 701–706, doi:10.1111/j.1523-1739.2005.00703.x.
  409. Schulz, J.J., L. Cayuela, C. Echeverria, J. Salas, and J.M. Rey Benayas, 2010: Monitoring land cover change of the Dryland forest landscape of Central Chile (1975–2008). Appl. Geogr., 30, 436–447, doi:10.1016/J.APGEOG.2009.12.003.
  410. Kassas, M., 1995: Desertification: A general review. J. Arid Environ., 30, 115–128, doi:10.1016/S0140-1963(05)80063-1.
  411. Ellis, J., 1995: Climate variability and complex ecosystem dynamics: Implications for pastoral development. In: Living with Uncertainty [Scoones I. (ed.)]. Practical Action Publishing, London, UK, pp. 37–46.
  412. Vetter, S., 2005: Rangelands at equilibrium and non-equilibrium: Recent developments in the debate. J. Arid Environ., 62, 321–341, doi:10.1016/J.JARIDENV.2004.11.015.
  413. von Wehrden, H., J. Hanspach, P. Kaczensky, J. Fischer, and K. Wesche, 2012: Global assessment of the non-equilibrium concept in rangelands. Ecol. Appl., 22, 393–399, doi:10.1890/11-0802.1.
  414. Prince, S. et al., 2018: Chapter 4: Status and trends of land degradation and restoration and associated changes in biodiversity and ecosystem functions. In: IPBES (2018): Assessment Report on Land Degradation and Restoration [Montanarella, L., R. Scholes., and A. Brainich. (eds.)]. Secretariat of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, Bonn, Germany, pp. 221–338.
  415. Prince, S.D., E.B. De Colstoun, and L.L. Kravitz, 1998: Evidence from rain-use efficiencies does not indicate extensive Sahelian desertification. Glob. Chang. Biol., 4, 359–374, doi:10.1046/j.1365-2486.1998.00158.x.
  416. Bai, Z.G., D.L. Dent, L. Olsson, and M.E. Schaepman, 2008: Proxy global assessment of land degradation. Soil Use Manag., 24, 223–234, doi:10.1111/j.1475-2743.2008.00169.x.
  417. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  418. Evans, J., and R. Geerken, 2004: Discrimination between climate and human-induced Dryland degradation. J. Arid Environ., 57, 535–554, doi:10.1016/S0140-1963(03)00121-6.
  419. Herrmann, S.M., and C.F. Hutchinson, 2005: The changing contexts of the desertification debate. J. Arid Environ., 63, 538–555, doi:10.1016/J.JARIDENV.2005.03.003.
  420. Huber, S., R. Fensholt, and K. Rasmussen, 2011: Water availability as the driver of vegetation dynamics in the African Sahel from 1982 to 2007. Glob. Planet. Change, 76, 186–195, doi:10.1016/J.GLOPLACHA.2011.01.006.
  421. Wessels, K.J., S.D. Prince, J. Malherbe, J. Small, P.E. Frost, and D. VanZyl, 2007: Can human-induced land degradation be distinguished from the effects of rainfall variability? A case study in South Africa. J. Arid Environ., 68, 271–297, doi:10.1016/J.JARIDENV.2006.05.015.
  422. Leroux, L., A. Bégué, D. Lo Seen, A. Jolivot, and F. Kayitakire, 2017: Driving forces of recent vegetation changes in the Sahel: Lessons learned from regional and local level analyses. Remote Sens. Environ., 191, 38–54, doi:10.1016/J.RSE.2017.01.014.
  423. Omuto, C.T., R.R. Vargas, M.S. Alim, and P. Paron, 2010: Mixed-effects modelling of time series NDVI-rainfall relationship for detecting human-induced loss of vegetation cover in drylands. J. Arid Environ., 74, 1552–1563, doi:10.1016/J.JARIDENV.2010.04.001.
  424. Ibrahim, Y., H. Balzter, J. Kaduk, and C. Tucker, 2015: Land degradation assessment using residual trend analysis of GIMMS NDVI3g, soil moisture and rainfall in Sub-Saharan West Africa from 1982 to 2012. Remote Sens., 7, 5471–5494, doi:10.3390/rs70505471.
  425. Li, Y. et al., 2012: Mongolian pine plantations enhance soil physico-chemical properties and carbon and nitrogen capacities in semi-arid degraded sandy land in China. Appl. Soil Ecol., 56, 1–9, doi:10.1016/J.APSOIL.2012.01.007.
  426. Yin, F., X. Deng, Q. Jin, Y. Yuan, and C. Zhao, 2014: The impacts of climate change and human activities on grassland productivity in Qinghai Province, China. Front. Earth Sci., 8, 93–103, doi:10.1007/s11707-013-0390-y.
  427. Jiang, L., G. Jiapaer, A. Bao, H. Guo, and F. Ndayisaba, 2017: Vegetation dynamics and responses to climate change and human activities in Central Asia. Sci. Total Environ., 599–600, 967–980, doi:10.1016/J.SCITOTENV.2017.05.012.
  428. Burrell, A.L., J.P. Evans, and Y. Liu, 2017: Detecting dryland degradation using Time Series Segmentation and Residual Trend analysis (TSS-RESTREND). Remote Sens. Environ., 197, doi:10.1016/j.rse.2017.05.018.
  429. Andela, N., Y.Y. Liu, A.I.J.M. van Dijk, R.A.M. de Jeu, and T.R. McVicar, 2013: Global changes in Dryland vegetation dynamics (1988-2008) assessed by satellite remote sensing: Comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences, 10, 6657–6676, doi:10.5194/bg-10-6657-2013.
  430. Liu, Y.Y. et al., 2013: Changing climate and overgrazing are decimating Mongolian steppes. PloS One, 8, e57599. 1-6, doi:10.1371/journal.pone.0057599.
  431. Higginbottom, T., and E. Symeonakis, 2014: Assessing land degradation 
and desertification using vegetation index data: Current frameworks and future directions. Remote Sens., 6, 9552–9575, doi:10.3390/rs6109552.
  432. Browning, D.M., J.J. Maynard, J.W. Karl, and D.C. Peters, 2017: Breaks in MODIS time series portend vegetation change: Verification using long-term data in an arid grassland ecosystem. Ecol. Appl., 27, 1677–1693, doi:10.1002/eap.1561.
  433. Burrell, A.L., J.P. Evans, and Y. Liu, 2017: Detecting dryland degradation using Time Series Segmentation and Residual Trend analysis (TSS-RESTREND). Remote Sens. Environ., 197, doi:10.1016/j.rse.2017.05.018.
  434. Horion, S., A.V. Prishchepov, J. Verbesselt, K. de Beurs, T. Tagesson, and 
R. Fensholt, 2016: Revealing turning points in ecosystem functioning over the northern Eurasian agricultural frontier. Glob. Chang. Biol., 22, 2801–2817, doi:10.1111/gcb.13267.
  435. Seaquist, J.W., T. Hickler, L. Eklundh, J. Ardö, and B.W. Heumann, 2009: Disentangling the effects of climate and people on Sahel vegetation dynamics. Biogeosciences, 6, 469–477, doi:10.5194/bg-6-469-2009.
  436. Mao, J., X. Shi, P. Thornton, F. Hoffman, Z. Zhu, and R. Myneni, 2013: Global latitudinal-asymmetric vegetation growth trends and their driving mechanisms: 1982–2009. Remote Sens., 5, 1484–1497, doi:10.3390/rs5031484.
  437. Zhu, Z. et al., 2016: Greening of the Earth and its drivers. Nat. Clim. Chang., 6, 791–795, doi:10.1038/nclimate3004.
  438. Piao, S. et al., 2015: Detection and attribution of vegetation greening trend in China over the last 30 years. Glob. Chang. Biol., 21, 1601–1609, doi:10.1111/gcb.12795.
  439. Mao, J. et al., 2016: Human-induced greening of the northern extratropical land surface. Nat. Clim. Chang., 6, 959–963, doi:10.1038/nclimate3056.
  440. Uhe, P. et al., 2018: Attributing drivers of the 2016 Kenyan drought. Int. J. Climatol., 38, e554–e568, doi:10.1002/joc.5389.
  441. Funk, C. et al., 2019: Examining the potential contributions of extreme ‘Western V’ sea surface temperatures to the 2017 March–June East African drought. Bull. Am. Meteorol. Soc., 100, S55–S60, doi:10.1175/BAMS-D-18-0108.1.
  442. Thomas, N., and S. Nigam, 2018: 20th century climate change over Africa: Seasonal hydroclimate trends and Sahara desert expansion. J. Clim., 31, 3349–3370, doi:10.1175/JCLI-D-17-0187.1.
  443. Masih, I., S. Maskey, F.E.F. Mussá, and P. Trambauer, 2014: A review of droughts on the African continent: A geospatial and long-term perspective. Hydrol. Earth Syst. Sci., 18, 3635–3649, doi:10.5194/hess-18-3635-2014.
  444. Kouba, Y., M. Gartzia, A. El Aich, and C.L. Alados, 2018: Deserts do not advance, they are created: Land degradation and desertification in semi-arid environments in the Middle Atlas, Morocco. J. Arid Environ., 158, 1–8, doi:10.1016/J.JARIDENV.2018.07.002.
  445. Inbar, M., 2007: Importance of drought information in monitoring and assessing land degradation. In: Climate and Land Degradation [Sivakumar, M.V.K. and N. Ndiang’ui (eds.)]. Springer Berlin Germany, pp. 253–266.
  446. Fu, B., L. Chen, K. Ma, H. Zhou, and J. Wang, 2000: The relationships between land use and soil conditions in the hilly area of the Loess Plateau in northern Shaanxi, China. CATENA, 39, 69–78, doi:10.1016/S0341-8162(99)00084-3.
  447. Liu, M., and H. Tian, 2010: China’s land cover and land use change from 1700 to 2005: Estimations from high-resolution satellite data and historical archives. Global Biogeochem. Cycles, 24, GB3003, 1-18, doi:10.1029/2009GB003687.
  448. Zhao, R., Y. Chen, P. Shi, L. Zhang, J. Pan, and H. Zhao, 2013: Land use and land cover change and driving mechanism in the arid inland river basin: A case study of Tarim River, Xinjiang, China. Environ. Earth Sci., 68, 591–604, doi:10.1007/s12665-012-1763-3.
  449. Kidron, G.J., and V.P. Gutschick, 2017: Temperature rise may explain grass depletion in the Chihuahuan Desert. Ecohydrology, 10, e1849, doi:10.1002/eco.1849.
  450. Lei, Y., B. Hoskins, and J. Slingo, 2011: Exploring the interplay between natural decadal variability and anthropogenic climate change in summer rainfall over China. Part I: Observational evidence. J. Clim., 24, 4584–4599, doi:10.1175/2010JCLI3794.1.
  451. Yang, Y., Z. Wang, J. Li, C. Gang, Y. Zhang, Y. Zhang, I. Odeh, and J. Qi, 2016: Comparative assessment of grassland degradation dynamics in response to climate variation and human activities in China, Mongolia, Pakistan and Uzbekistan from 2000 to 2013. J. Arid Environ., 135, 164–172, doi:10.1016/J.JARIDENV.2016.09.004.
  452. Rasmussen, K., S. D’haen, R. Fensholt, B. Fog, S. Horion, J.O. Nielsen, 
L.V. Rasmussen, and A. Reenberg, 2016: Environmental change in the Sahel: Reconciling contrasting evidence and interpretations. Reg. Environ. Chang., 16, 673–680, doi:10.1007/s10113-015-0778-1.
  453. Asner, G.P. et al., 2012: Carnegie Airborne Observatory-2: Increasing science data dimensionality via high-fidelity multi-sensor fusion. Remote Sens. Environ., 124, 454–465, doi:10.1016/J.RSE.2012.06.012.
  454. Sivakumar, M.V.K., Ndiang’ui, N. (ed.), 2007: Climate and Land Degradation (Environmental Science and Engineering). Springer-Verlag, Berlin, Germany, 623 pp.
  455. Wang, K., and R.E. Dickinson, 2013: Contribution of solar radiation to decadal temperature variability over land. Proc. Natl. Acad. Sci. U.S.A., 110, 14877–14882, doi:10.1073/pnas.1311433110.
  456. Seneviratne, S.I., T. Corti, E.L. Davin, M. Hirschi, E.B. Jaeger, I. Lehner, 
B. Orlowsky, and A.J. Teuling, 2010: Investigating soil moisture-climate interactions in a changing climate: A review. Earth-Science Rev., 99, 125–161, doi:10.1016/J.EARSCIREV.2010.02.004.
  457. UNEP WMO UNCCD, 2016: Global Assessment of Sand and Dust Storms. United Nations Environment Programme, Nairobi, Kenya, 139 pp.
  458. Middleton, N., and U. Kang, 2017: Sand and dust storms: Impact mitigation. Sustainability, 9, 1053, doi:10.3390/su9061053.
  459. Ginoux, P., J.M. Prospero, T.E. Gill, N.C. Hsu, and M. Zhao, 2012: Global-scale attribution of anthropogenic and natural dust sources and their emission rates based on MODIS Deep Blue aerosol products. Rev. Geophys., 50, doi:10.1029/2012RG000388.
  460. Rajot, J.L., S.C. Alfaro, L. Gomes, and A. Gaudichet, 2003: Soil crusting on sandy soils and its influence on wind erosion. CATENA, 53, 1–16, doi:10.1016/S0341-8162(02)00201-1.
  461. Field, J.P. et al., 2010: The ecology of dust. Front. Ecol. Environ., 8, 423–430, doi:10.1890/090050.
  462. Rodriguez-Caballero, E., J. Belnap, B. Büdel, P.J. Crutzen, M.O. Andreae, U. Pöschl, and B. Weber, 2018: Dryland photoautotrophic soil surface communities endangered by global change. Nat. Geosci., 11, 185–189, doi:10.1038/s41561-018-0072-1.
  463. Choobari, O.A., P. Zawar-Reza, and A. Sturman, 2014: The global distribution of mineral dust and its impacts on the climate system: A review. Atmos. Res., 138, 152–165, doi:10.1016/J.ATMOSRES.2013.11.007.
  464. Kaufman, Y.J., D. Tanré, and O. Boucher, 2002: A satellite view of aerosols in the climate system. Nature, 419, 215–223, doi:10.1038/nature01091.
  465. Middleton, N., and U. Kang, 2017: Sand and dust storms: Impact mitigation. Sustainability, 9, 1053, doi:10.3390/su9061053.
  466. Kok, J.F., D.S. Ward, N.M. Mahowald, and A.T. Evan, 2018: Global and regional importance of the direct dust-climate feedback. Nat. Commun., 9, 241, doi:10.1038/s41467-017-02620-y.
  467. Huang, J., H. Yu, A. Dai, Y. Wei, and L. Kang, 2017: Drylands face potential threat under 2°C global warming target. Nat. Clim. Chang., 7, 417–422, doi:10.1038/nclimate3275.
  468. Kaufman, Y.J., D. Tanré, and O. Boucher, 2002: A satellite view of aerosols in the climate system. Nature, 419, 215–223, doi:10.1038/nature01091.
  469. Boucher, O., D. Randall, P. Artaxo, C. Bretherton, G. Feingold, P. Forster, 
V.-M. Kerminen, Y. Kondo, H. Liao, U. Lohmann, P. Rasch, S.K. Satheesh, 
S. Sherwood, B. Stevens and X.Y. Zhang, 2013: Clouds and aerosols. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, 
J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdon and New York, NY, USA, pp. 571–657.
  470. Engelstaedter, S., K.E. Kohfeld, I. Tegen, and S.P. Harrison, 2003: Controls of dust emissions by vegetation and topographic depressions: An evaluation using dust storm frequency data. Geophys. Res. Lett., 30 
1294, doi:10.1029/2002GL016471.
  471. Fan, B. et al., 2015: Earlier vegetation green-up has reduced spring dust storms. Sci. Rep., 4, 6749, doi:10.1038/srep06749.
  472. Yu, K., P. D’Odorico, A. Bhattachan, G.S. Okin, and A.T. Evan, 2015: Dust-rainfall feedback in West African Sahel. Geophys. Res. Lett., 42, 7563–7571, doi:10.1002/2015GL065533.
  473. Zou, X.K., and P.M. Zhai, 2004: Relationship between vegetation coverage and spring dust storms over northern China. J. Geophys. Res. Atmos., 109, D03104: 1-9, doi:10.1029/2003JD003913.
  474. Elmore, A.J., J.M. Kaste, G.S. Okin, and M.S. Fantle, 2008: Groundwater influences on atmospheric dust generation in deserts. J. Arid Environ., 72, 1753–1765, doi:10.1016/J.JARIDENV.2008.05.008.
  475. Decker, M., A.J. Pitman, J.P. Evans, M. Decker, A.J. Pitman, and J.P. Evans, 2013: Groundwater constraints on simulated transpiration variability over south-eastern Australian forests. J. Hydrometeorol., 14, 543–559, doi:10.1175/JHM-D-12-058.1.
  476. Lamontagne, S., P.G. Cook, A. O’Grady, and D. Eamus, 2005: Groundwater use by vegetation in a tropical savanna riparian zone (Daly River, Australia). J. Hydrol., 310, 280–293, doi:10.1016/J.JHYDROL.2005.01.009.
  477. Patten, D.T., L. Rouse, and J.C. Stromberg, 2008: Isolated spring wetlands in the Great Basin and Mojave Deserts, USA: Potential response of vegetation to groundwater withdrawal. Environ. Manage., 41, 398–413, doi:10.1007/s00267-007-9035-9.
  478. Cook, B.I., R.L. Miller, and R. Seager, 2009: Amplification of the North American ‘Dust Bowl’ drought through human-induced land degradation. Proc. Natl. Acad. Sci. U.S.A., 106, 4997–5001, doi:10.1073/pnas.0810200106.
  479. Huang, J., Q. Fu, J. Su, Q. Tang, P. Minnis, Y. Hu, Y. Yi, and Q. Zhao, 2009: Taklimakan dust aerosol radiative heating derived from CALIPSO observations using the Fu-Liou radiation model with CERES constraints. Atmos. Chem. Phys., 9, 4011–4021, doi:10.5194/acp-9-4011-2009.
  480. Konare, A. et al., 2008: A regional climate modeling study of the effect of desert dust on the West African monsoon. J. Geophys. Res., 113, D12206, doi:10.1029/2007JD009322.
  481. Rosenfeld, D., Y. Rudich, and R. Lahav, 2001: Desert dust suppressing precipitation: A possible desertification feedback loop. Proc. Natl. Acad. Sci. U.S.A., 98, 5975–5980, doi:10.1073/pnas.101122798.
  482. Solmon, F., N. Elguindi, and M. Mallet, 2012: Radiative and climatic effects of dust over West Africa, as simulated by a regional climate model. Clim. Res., 52, 97–113, doi:10.3354/cr01039.
  483. Zhao, S., H. Zhang, S. Feng, and Q. Fu, 2015: Simulating direct effects of dust aerosol on arid and semi-arid regions using an aerosol-climate coupled system. Int. J. Climatol., 35, 1858–1866, doi:10.1002/joc.4093.
  484. Islam, M.N., and M. Almazroui, 2012: Direct effects and feedback of desert dust on the climate of the Arabian Peninsula during the wet season: 
A regional climate model study. Clim. Dyn., 39, 2239–2250, doi:10.1007/s00382-012-1293-4.
  485. Lau, K.M., K.M. Kim, Y.C. Sud, and G.K. Walker, 2009: A GCM study of the response of the atmospheric water cycle of West Africa and the Atlantic to Saharan dust radiative forcing. Ann. Geophys., 27, 4023–4037, doi:10.5194/angeo-27-4023-2009.
  486. Sun, H., Z. Pan, and X. Liu, 2012: Numerical simulation of spatial-temporal distribution of dust aerosol and its direct radiative effects on East Asian climate. J. Geophys. Res. Atmos., 117, D13206: 1-14, doi:10.1029/2011JD017219.
  487. Huang, J., T. Wang, W. Wang, Z. Li, and H. Yan, 2014: Climate effects of dust aerosols over East Asian arid and semi-arid regions. J. Geophys. Res. Atmos., 119, 11, 311–398, 416, doi:10.1002/2014JD021796.
  488. Miller, R.L., I. Tegen, and J. Perlwitz, 2004: Surface radiative forcing by soil dust aerosols and the hydrologic cycle. J. Geophys. Res. Atmos., 109, D04203: 1–24 D04203,n/a, doi:10.1029/2003JD004085.
  489. Zhao, S., H. Zhang, S. Feng, and Q. Fu, 2015: Simulating direct effects of dust aerosol on arid and semi-arid regions using an aerosol-climate coupled system. Int. J. Climatol., 35, 1858–1866, doi:10.1002/joc.4093.
  490. Jickells, T.D. et al., 2005: Global iron connections between desert dust, ocean biogeochemistry, and climate. Science, 308, 67–71, doi:10.1126/science.1105959.
  491. Okin, G.S. et al., 2011: Impacts of atmospheric nutrient deposition on marine productivity: Roles of nitrogen, phosphorus, and iron. Global Biogeochem. Cycles, 25, n/a–n/a, doi:10.1029/2010GB003858.
  492. Das, R., A. Evan, and D. Lawrence, 2013: Contributions of long-distance dust transport to atmospheric P inputs in the Yucatan Peninsula. Global Biogeochem. Cycles, 27, 167–175, doi:10.1029/2012GB004420.
  493. Kaufman, Y.J. et al., 2005: Dust transport and deposition observed from the Terra-Moderate Resolution Imaging Spectroradiometer (MODIS) spacecraft over the Atlantic Ocean. J. Geophys. Res., 110, D10S12, doi:10.1029/2003JD004436.
  494. Uno, I. et al., 2009: Asian dust transported one full circuit around the globe. Nat. Geosci., 2, 557–560, doi:10.1038/ngeo583.
  495. Zhang, X.Y., R. Arimoto, and Z.S. An, 1997: Dust emission from Chinese desert sources linked to variations in atmospheric circulation. J. Geophys. Res. Atmos., 102, 28041–28047, doi:10.1029/97JD02300.
  496. Varga, G., G. Újvári, and J. Kovács, 2014: Spatiotemporal patterns of Saharan dust outbreaks in the Mediterranean Basin. Aeolian Res., 15, 151–160, doi:10.1016/j.aeolia.2014.06.005.
  497. Evan, A.T., and S. Mukhopadhyay, 2010: African Dust over the Northern Tropical Atlantic: 1955–2008. J. Appl. Meteorol. Climatol., 49, 2213–2229, doi:10.1175/2010JAMC2485.1.
  498. Evan, A.T., D.J. Vimont, A.K. Heidinger, J.P. Kossin, and R. Bennartz, 2009: The role of aerosols in the evolution of tropical North Atlantic Ocean temperature anomalies. Science, 324, 778–781, doi:10.1126/science.1167404.
  499. Evan, A.T., D.J. Vimont, A.K. Heidinger, J.P. Kossin, and R. Bennartz, 2009: The role of aerosols in the evolution of tropical North Atlantic Ocean temperature anomalies. Science, 324, 778–781, doi:10.1126/science.1167404.
  500. Lenes, J.M. et al., 2001: Iron fertilization and the Trichodesmium response on the West Florida shelf. Limnol. Oceanogr., 46, 1261–1277, doi:10.4319/lo.2001.46.6.1261.
  501. Shaw, E.C., A.J. Gabric, and G.H. McTainsh, 2008: Impacts of aeolian dust deposition on phytoplankton dynamics in Queensland coastal waters. Mar. Freshw. Res., 59, 951, doi:10.1071/MF08087.
  502. Neuer, S., M.E. Torres-Padrón, M.D. Gelado-Caballero, M.J. Rueda, 
J. Hernández-Brito, R. Davenport, and G. Wefer, 2004: Dust deposition pulses to the eastern subtropical North Atlantic gyre: Does ocean’s biogeochemistry respond? Global Biogeochem. Cycles, 18, GB4020, doi:10.1029/2004GB002228.
  503. Boyd, P.W., D.S. Mackie, and K.A. Hunter, 2010: Aerosol iron deposition to the surface ocean – Modes of iron supply and biological responses. Mar. Chem., 120, 128–143, doi:10.1016/J.MARCHEM.2009.01.008.
  504. Painter, T.H., S.M. Skiles, J.S. Deems, W.T. Brandt, and J. Dozier, 2018: Variation in rising limb of Colorado River snowmelt runoff hydrograph controlled by dust radiative forcing in snow. Geophys. Res. Lett., 45, 797–808, doi:10.1002/2017GL075826.
  505. Kaspari, S., T.H. Painter, M. Gysel, S.M. Skiles, and M. Schwikowski, 2014: Seasonal and elevational variations of black carbon and dust in snow and ice in the Solu-Khumbu, Nepal and estimated radiative forcings. Atmos. Chem. Phys., 14, 8089–8103, doi:10.5194/acp-14-8089-2014.
  506. Qian, Y. et al., 2015: Light-absorbing particles in snow and ice: Measurement and modeling of climatic and hydrological impact. Adv. Atmos. Sci., 32, 64–91, doi:10.1007/s00376-014-0010-0.
  507. Painter, T.H., M.G. Flanner, G. Kaser, B. Marzeion, R.A. VanCuren, and 
W. Abdalati, 2013: End of the Little Ice Age in the Alps forced by industrial black carbon. Proc. Natl. Acad. Sci., 110, 15216–15221, doi:10.1073/PNAS.1302570110.
  508. Charney, J., P.H. Stone, and W.J. Quirk, 1975: Drought in the Sahara: A biogeophysical feedback mechanism. Science, 187, 434–435, doi:10.1126/science.187.4175.434.
  509. Zeng, N., and J. Yoon, 2009: Expansion of the world’s deserts due to vegetation-albedo feedback under global warming. Geophys. Res. Lett., 36, L17401, doi:10.1029/2009GL039699.
  510. Zaitchik, B.F. et al., 2007: Climate and vegetation in the Middle East: Interannual variability and drought feedbacks. J. Clim., 20, 3924–3941, doi:10.1175/JCLI4223.1.
  511. Evans, J.P., X. Meng, and M.F. McCabe, 2017: Land surface albedo and vegetation feedbacks enhanced the millennium drought in Southeast Australia. Hydrol. Earth Syst. Sci., 21, 409–422, doi:10.5194/hess-21-409-2017.
  512. Meng, X.H., J.P. Evans, and M.F. McCabe, 2014a: The influence of inter-annually varying albedo on regional climate and drought. Clim. Dyn., 42, 787–803, doi:10.1007/s00382-013-1790-0.
  513. Meng X.H., J.P. Evans, and M.F. McCabe, 2014b: The impact of observed vegetation changes on land-atmosphere feedbacks during drought. J. Hydrometeorol., 15, 759–776, doi:10.1175/JHM-D-13-0130.1.
  514. Lee, S.-J., and E.H. Berbery, 2012: Land cover change effects on the climate of the La Plata Basin. J. Hydrometeorol., 13, 84–102, doi:10.1175/JHM-D-11-021.1.
  515. Zaitchik, B.F. et al., 2013: Representation of soil moisture feedbacks during drought in NASA Unified WRF (NU-WRF). J. Hydrometeorol., 14, 360–367, doi:10.1175/JHM-D-12-069.1.
  516. Rodriguez-Caballero, E., J. Belnap, B. Büdel, P.J. Crutzen, M.O. Andreae, U. Pöschl, and B. Weber, 2018: Dryland photoautotrophic soil surface communities endangered by global change. Nat. Geosci., 11, 185–189, doi:10.1038/s41561-018-0072-1.
  517. Rutherford, W.A., T.H. Painter, S. Ferrenberg, J. Belnap, G.S. Okin, C. Flagg, and S.C. Reed, 2017: Albedo feedbacks to future climate via climate change impacts on dryland biocrusts. Sci. Rep., 7, 44188, doi:10.1038/srep44188.
  518. Rotenberg, E. and D. Yakir, 2010: Contribution of semi-arid forests to the climate system. Science, 327, 451–454, doi:10.1126/science.1179998.
  519. Schlesinger, W.H. et al., 1990: Biological feedbacks in global desertification. Science, 4946, 1043–1048, doi:10.1126/science.247.4946.1043.
  520. Sivakumar, M.V.K., Ndiang’ui, N. (ed.), 2007: Climate and Land Degradation (Environmental Science and Engineering). Springer-Verlag, Berlin, Germany, 623 pp.
  521. Sivakumar, M.V.K., 2007: Interactions between climate and desertification. Agric. For. Meteorol., 142, 143–155, doi:10.1016/j.agrformet.2006.03.025.
  522. Plaza, C. et al., 2018: Soil resources and element stocks in drylands to face global issues. Sci. Rep., 8, 13788, doi:10.1038/s41598-018-32229-0.
  523. Abril, A., P. Barttfeld, and E.H. Bucher, 2005: The effect of fire and overgrazing disturbes on soil carbon balance in the Dry Chaco forest. For. Ecol. Manage., 206, 399–405, doi:10.1016/j.foreco.2004.11.014. doi.org/10.1016/j.foreco.2004.
11.014.
  524. Asner, G.P., S. Archer, R.F. Hughes, R.J. Ansley, and C.A. Wessman, 2003: Net changes in regional woody vegetation cover and carbon storage in Texas drylands, 1937–1999. Glob. Chang. Biol., 9, 316–335, doi:10.1046/j.1365-2486.2003.00594.x.
  525. Abdalla, M. et al., 2018: Critical review of the impacts of grazing intensity on soil organic carbon storage and other soil quality indicators in extensively managed grasslands. Agric. Ecosyst. Environ., 253, 62–81, doi:10.1016/j.agee.2017.10.023.
  526. Lechmere-Oertel, R.G., G.I.H. Kerley, and R.M. Cowling, 2005: Patterns and implications of transformation in semi-arid succulent thicket, South Africa. J. Arid Environ., 62, 459–474, doi:10.1016/j.jaridenv.2004.11.016.
  527. Maestre, F.T. et al., 2012: Plant species richness and ecosystems multifunctionality in global drylands. Science, 335, 2014–2017, doi:10.1126/science.1215442.
  528. Lechmere-Oertel, R.G., G.I.H. Kerley, and R.M. Cowling, 2005: Patterns and implications of transformation in semi-arid succulent thicket, South Africa. J. Arid Environ., 62, 459–474, doi:10.1016/j.jaridenv.2004.11.016.
  529. Powell, M.J., 2009: Restoration of Degraded Subtropical Thickets in the Baviaanskloof Megareserve, South Africa. Master Thesis, Rhodes University, Grahamstown, South Africa.
  530. Scholes, R.J., and B.H. Walker and B.H. Walker, 1993: An African Savanna: Synthesis of the Nylsvley Study. Cambridge University Press, Cambridge, UK. https://doi.org/10.1017/CBO9780511565472, 306 pp.
  531. Woomer, P.L., A. Touré, and M. Sall, 2004: Carbon stocks in Senegal’s Sahel transition zone. J. Arid Environ., 59, 499–510. doi.org/10.1016/j.jaridenv.2004.03.027.
  532. Asner, G.P., S. Archer, R.F. Hughes, R.J. Ansley, and C.A. Wessman, 2003: Net changes in regional woody vegetation cover and carbon storage in Texas drylands, 1937–1999. Glob. Chang. Biol., 9, 316–335, doi:10.1046/j.1365-2486.2003.00594.x.
  533. Hughes, R.F. et al., 2006: Changes in aboveground primary production and carbon and nitrogen pools accompanying woody plant encroachment in a temperate savanna. Glob. Chang. Biol., 12, 1733–1747, doi:10.1111/j.1365-2486.2006.01210.x.
  534. Petrie, M.D., S.L. Collins, A.M. Swann, P.L. Ford, and M.E. Litvak, 2015: Grassland to shrubland state transitions enhance carbon sequestration in the northern Chihuahuan Desert. Glob. Chang. Biol., 21, 1226–1235, doi:10.1111/gcb.12743.
  535. Li, H. et al., 2016: Effects of shrub encroachment on soil organic carbon in global grasslands. Sci. Rep., 6, 28974, doi:10.1038/srep28974.
  536. Asner, G.P., S. Archer, R.F. Hughes, R.J. Ansley, and C.A. Wessman, 2003: Net changes in regional woody vegetation cover and carbon storage in Texas drylands, 1937–1999. Glob. Chang. Biol., 9, 316–335, doi:10.1046/j.1365-2486.2003.00594.x.
  537. Hudak, A.T., C.A. Wessman, and T.R. Seastedt, 2003: Woody overstorey effects on soil carbon and nitrogen pools in a South African savanna. Austral Ecol., 28, 173–181. doi:10.1046/j.1442-9993.2003.01265.x.
  538. Barger, N.N. et al., 2011: Woody plant proliferation in North American drylands: A synthesis of impacts on ecosystem carbon balance. J. Geophys. Res. Biogeosciences, 116:1-17, G00K07, doi:10.1029/2010JG001506.
  539. Jackson, R.B., J.L. Banner, E.G. Jobbagy, W.T. Pockman, and D.H. Wall, 2002: Ecosystem carbon loss with woody plant invasion of grasslands. Nature, 418, 623–626, doi:10.1038/nature00910.
  540. Magandana, T.P., 2016: Effect of Acacia karoo encroachment on grass production in the semi-arid savannas of the Eastern Cape, South Africa. Thesis, University of Fort Hare, Alice, South Africa, 104 pp.
  541. Archibald, S., D.P. Roy, B.W. Van Wilgen, and R.J. Scholes, 2009: What limits fire? An examination of drivers of burnt area in southern Africa. Glob. Chang. Biol., 15, 613–630, doi:10.1111/j.1365-2486.2008.01754.x.
  542. Millennium Ecosystem Assessment, 2005: Ecosystems and Human Well-Being: Desertification Synthesis. World Resource Institute, Washington, DC, 
USA, 26 pp.
  543. Enfors, E.I., and L.J. Gordon, 2008: Dealing with drought: The challenge of using water system technologies to break Dryland poverty traps. Glob. Environ. Chang., 18, 607–616, doi:10.1016/J.GLOENVCHA.2008.07.006.
  544. Mortimore, M., 2005: Dryland development: Success stories from West Africa. Environment, 47, 8–21, doi:10.3200/ENVT.47.1.8-21.
  545. Majeed, A. and Z. Muhammad, 2019: Salinity: A major agricultural problem – Causes, impacts on crop productivity and management strategies. In: Plant Abiotic Stress Tolerance [Hasanuzzaman, M., K.R. Hakeem, K. Nahar and H. Alharby (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 83–99.
  546. Mirzabaev, A., M. Ahmed, J. Werner, J. Pender, and M. Louhaichi, 2016a: Rangelands of Central Asia: Challenges and opportunities. J. Arid Land, 8, 93–108, doi:10.1007/s40333-015-0057-5.
  547. Qadir, M., A.D. Noble, A.S. Qureshi, R.K. Gupta, T. Yuldashev, and A. Karimov, 2009: Salt-induced land and water degradation in the Aral Sea Basin: A challenge to sustainable agriculture in Central Asia. Nat. Resour. Forum, 33, 134–149, doi:10.1111/j.1477-8947.2009.01217.x.
  548. Van Loo, M. et al., 2017: Human induced soil erosion and the implications on crop yield in a small mountainous Mediterranean catchment (SW-Turkey). CATENA, 149, 491–504, doi:10.1016/J.CATENA.2016.08.023.
  549. Tokbergenova, A., G. Nyussupova, M. Arslan, and S.K.L. Kiyassova, 2018: Causes and impacts of land degradation and desertification: Case study from Kazakhstan. In: Vegetation of Central Asia and Environs [Egamberdieva, D. and M. Ozturk (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 291–302.
  550. Zika, M., and K.H. Erb, 2009: The global loss of net primary production resulting from human-induced soil degradation in drylands. Ecol. Econ., 69, 310–318, doi:10.1016/j.ecolecon.2009.06.014.
  551. Ivanov, A.L. et al., 2018: National report global climate and soil cover of Russia: Assessment of risks and environmental and economic consequences of land degradation. Adaptive systems and technologies of environmental management (agriculture and forestry). In: Russian: NATSIONAL’N. Moscow, Russia, 357 pp.
  552. Ivanov, A.L. et al., 2018: National report global climate and soil cover of Russia: Assessment of risks and environmental and economic consequences of land degradation. Adaptive systems and technologies of environmental management (agriculture and forestry). In: Russian: NATSIONAL’N. Moscow, Russia, 357 pp.
  553. Rojas-Downing, M.M., A.P. Nejadhashemi, T. Harrigan, and S.A. Woznicki, 2017: Climate change and livestock: Impacts, adaptation, and mitigation. Clim. Risk Manag., 16, 145–163, doi:10.1016/j.crm.2017.02.001.
  554. Nardone, A., B. Ronchi, N. Lacetera, M.S. Ranieri, and U. Bernabucci, 2010: Effects of climate changes on animal production and sustainability of livestock systems. Livest. Sci., 130, 57–69, doi:10.1016/j.livsci.2010.02.011.
  555. Thornton, P.K., J. van de Steeg, A. Notenbaert, and M. Herrero, 2009: The impacts of climate change on livestock and livestock systems in developing countries: A review of what we know and what we need to know. Agric. Syst., 101, 113–127, doi:10.1016/J.AGSY.2009.05.002.
  556. Nardone, A., B. Ronchi, N. Lacetera, M.S. Ranieri, and U. Bernabucci, 2010: Effects of climate changes on animal production and sustainability of livestock systems. Livest. Sci., 130, 57–69, doi:10.1016/j.livsci.2010.02.011.
  557. Thornton, P.K., J. van de Steeg, A. Notenbaert, and M. Herrero, 2009: The impacts of climate change on livestock and livestock systems in developing countries: A review of what we know and what we need to know. Agric. Syst., 101, 113–127, doi:10.1016/J.AGSY.2009.05.002.
  558. Polley, H.W. et al., 2013: Climate change and North American rangelands: Trends, projections, and implications. Rangel. Ecol. Manag., 66, 493–511, doi:10.2111/REM-D-12-00068.1.
  559. Mudd, G.M., 2000: Mound springs of the Great Artesian Basin in South Australia: A case study from Olympic Dam. Environ. Geol., 39, 463–476, doi:10.1007/s002540050452.
  560. Mays, L.W., 2013: Groundwater resources sustainability: Past, present, and future. Water Resour. Manag., 27, 4409–4424, doi:10.1007/s11269-013-0436-7.
  561. Mahmod, W.E., and K. Watanabe, 2014: Modified Grey Model and its application to groundwater flow analysis with limited hydrogeological data: A case study of the Nubian Sandstone, Kharga Oasis, Egypt. Environ. Monit. Assess., 186, 1063–1081, doi:10.1007/s10661-013-3439-1.
  562. Jolly, I.D., K.L. Mcewan, and K.L. Holland, 2008: A review of groundwater – Surface water interactions in arid/semi-arid wetlands and the consequences of salinity for wetland ecology. Ecohydrology, 1, 43–58, doi:10.1002/eco.6.
  563. Konikow, L.F., 2011: Contribution of global groundwater depletion since 1900 to sea-level rise. Geophys. Res. Lett., 38, 1–5, doi:10.1029/2011GL048604.
  564. Squeo, F.A., R. Aravena, E. Aguirre, A. Pollastri, C.B. Jorquera, and J.R. Ehleringer, 2006: Groundwater dynamics in a coastal aquifer in north-central Chile: Implications for groundwater recharge in an arid ecosystem. J. Arid Environ., 67, 240–254, doi:10.1016/j.jaridenv.2006.02.012.
  565. Mahmod, W.E., and K. Watanabe, 2014: Modified Grey Model and its application to groundwater flow analysis with limited hydrogeological data: A case study of the Nubian Sandstone, Kharga Oasis, Egypt. Environ. Monit. Assess., 186, 1063–1081, doi:10.1007/s10661-013-3439-1.
  566. Herrera, C. et al., 2018: Science of the total environment groundwater origin and recharge in the hyper-arid Cordillera de la Costa, Atacama Desert, northern Chile. Sci. Total Environ., 624, 114–132, doi:10.1016/j.scitotenv.2017.12.134.
  567. FAO, 1995: Desertification and Drought – Extent and Consequences Proposal for a Participatory Approach to Combat Desertification. Proceedings of a conference, Food and Agriculture Organization of the United Nations, Rome, Italy, 15 pp.
  568. Albaladejo, J., R. Ortiz, N. Garcia-Franco, A.R. Navarro, M. Almagro, J.G. Pintado, and M. Martínez-Mena, 2013: Land use and climate change impacts on soil organic carbon stocks in semi-arid Spain. J. Soils Sediments, 13, 265–277, doi:10.1007/s11368-012-0617-7.
  569. Almagro, M., J. López, C. Boix-Fayos, J. Albaladejo, and M. Martínez-Mena, 2010: Below-ground carbon allocation patterns in a dry Mediterranean ecosystem: A comparison of two models. Soil Biol. Biochem., 42, 1549–1557, doi:10.1016/j.soilbio.2010.05.031.
  570. Hoffmann, U., A. Yair, H. Hikel, and N.J. Kuhn, 2012: Soil organic carbon in the rocky desert of northern Negev (Israel). J. Soils Sediments, 12, 811–825, doi:10.1007/s11368-012-0499-8.
  571. Lavee, H., A.C. Imeson, and P. Sarah, 1998: The impact of climate change on geomorphology and desertification along a Mediterranean arid transect. L. Degrad. Dev., 9, 407–422, doi:10.1002/(SICI)1099-145X(199809/10)9:5<407::AID-LDR302>3.0.CO;2-6.
  572. Rey, A., E. Pegoraro, C. Oyonarte, A. Were, P. Escribano, and J. Raimundo, 2011: Impact of land degradation on soil respiration in a steppe (Stipa tenacissima L.) semi-arid ecosystem in the SE of Spain. Soil Biol. Biochem., 43, 393–403, doi:10.1016/j.soilbio.2010.11.007.
  573. Lavee, H., A.C. Imeson, and P. Sarah, 1998: The impact of climate change on geomorphology and desertification along a Mediterranean arid transect. L. Degrad. Dev., 9, 407–422, doi:10.1002/(SICI)1099-145X(199809/10)9:5<407::AID-LDR302>3.0.CO;2-6.
  574. Martinez-Mena, M., J. Lopez, M. Almagro, C. Boix-Fayos, and J. Albaladejo, 2008: Effect of water erosion and cultivation on the soil carbon stock in a semi-arid area of Southeast Spain. Soil Tillage Res., 99, 119–129, doi:10.1016/j.still.2008.01.009.
  575. Pierzynski, G., Brajendra, L. Caon, and R. Vargas, 2017: Threats to Soils: Global Trends and Perspectives. Global Land Outlook Working Paper 28. UNCCD, Bonn, Germany, 27 pp.
  576. Fay, P.A., D.M. Kaufman, J.B. Nippert, J.D. Carlisle, and C.W. Harper, 2008: Changes in grassland ecosystem function due to extreme rainfall events: Implications for responses to climate change. Glob. Chang. Biol., 14, 1600–1608, doi:10.1111/j.1365-2486.2008.01605.x.
  577. Hao, Y. et al., 2008: CO2, H2O and energy exchange of an Inner Mongolia steppe ecosystem during a dry and wet year. Acta Oecologica, 33, 133–143, doi:10.1016/j.actao.2007.07.002.
  578. Mi, J., J. Li, D. Chen, Y. Xie, and Y. Bai, 2015: Predominant control of moisture on soil organic carbon mineralization across a broad range of arid and semi-arid ecosystems on the Mongolia Plateau. Landsc. Ecol., 30, 1683–1699, doi:10.1007/s10980-014-0040-0.
  579. Serrano-Ortiz, P. et al., 2015: Seasonality of net carbon exchanges of Mediterranean ecosystems across an altitudinal gradient. J. Arid Environ., 115, 1–9, doi:10.1016/j.jaridenv.2014.12.003.
  580. Vargas, R. et al., 2012: Precipitation variability and fire influence the temporal dynamics of soil CO2 efflux in an arid grassland. Glob. Chang. Biol., 18, 1401–1411, doi:10.1111/j.1365-2486.2011.02628.x.
  581. Sharkhuu, A., A.F. Plante, O. Enkhmandal, C. Gonneau, B.B. Casper, B. Boldgiv, and P.S. Petraitis, 2016: Soil and ecosystem respiration responses to grazing, watering and experimental warming chamber treatments across topographical gradients in northern Mongolia. Geoderma, 269, 91–98, doi:10.1016/j.geoderma.2016.01.041.
  582. Biederman, J.A. et al., 2017: CO2 exchange and evapotranspiration across Dryland ecosystems of south-western North America. Glob. Chang. Biol., 23, 4204–4221, doi:10.1111/gcb.13686.
  583. Chen, S., G. Lin, J. H, and G.D. Jenerette, 2009: Dependence of carbon sequestration on the differential responses of ecosystem photosynthesis and respiration to rain pulses in a semi-arid steppe. Glob. Chang. Biol., 15, 2450–2461, doi:10.1111/j.1365-2486.2009.01879.x.
  584. Fay, P.A., D.M. Kaufman, J.B. Nippert, J.D. Carlisle, and C.W. Harper, 2008: Changes in grassland ecosystem function due to extreme rainfall events: Implications for responses to climate change. Glob. Chang. Biol., 14, 1600–1608, doi:10.1111/j.1365-2486.2008.01605.x.
  585. Hao, Y. et al., 2008: CO2, H2O and energy exchange of an Inner Mongolia steppe ecosystem during a dry and wet year. Acta Oecologica, 33, 133–143, doi:10.1016/j.actao.2007.07.002.
  586. Mi, J., J. Li, D. Chen, Y. Xie, and Y. Bai, 2015: Predominant control of moisture on soil organic carbon mineralization across a broad range of arid and semi-arid ecosystems on the Mongolia Plateau. Landsc. Ecol., 30, 1683–1699, doi:10.1007/s10980-014-0040-0.
  587. Sharkhuu, A., A.F. Plante, O. Enkhmandal, C. Gonneau, B.B. Casper, B. Boldgiv, and P.S. Petraitis, 2016: Soil and ecosystem respiration responses to grazing, watering and experimental warming chamber treatments across topographical gradients in northern Mongolia. Geoderma, 269, 91–98, doi:10.1016/j.geoderma.2016.01.041.
  588. Austin, A.T. et al., 2004: Water pulses and biogeochemical cycles in arid and semi-arid ecosystems. Oecologia, 141, 221–235, doi:10.1007/s00442-004-1519-1.
  589. Rutledge, S., D.I. Campbell, D. Baldocchi, and L.A. Schipper, 2010: Photodegradation leads to increased carbon dioxide losses from terrestrial organic matter. Glob. Chang. Biol., 16, 3065–3074, doi:10.1111/j.1365-2486.2009.02149.x.
  590. Biederman, J.A. et al., 2017: CO2 exchange and evapotranspiration across Dryland ecosystems of south-western North America. Glob. Chang. Biol., 23, 4204–4221, doi:10.1111/gcb.13686.
  591. Chen, S., G. Lin, J. H, and G.D. Jenerette, 2009: Dependence of carbon sequestration on the differential responses of ecosystem photosynthesis and respiration to rain pulses in a semi-arid steppe. Glob. Chang. Biol., 15, 2450–2461, doi:10.1111/j.1365-2486.2009.01879.x.
  592. Hao, Y. et al., 2008: CO2, H2O and energy exchange of an Inner Mongolia steppe ecosystem during a dry and wet year. Acta Oecologica, 33, 133–143, doi:10.1016/j.actao.2007.07.002.
  593. Poulter, B. et al., 2014: Contribution of semi-arid ecosystems to interannual variability of the global carbon cycle. Nature, 509, 600–603, doi:10.1038/nature13376.
  594. Donat, M.G., A.L. Lowry, L.V. Alexander, P.A. O’Gorman, and N. Maher, 2016: More extreme precipitation in the world’s dry and wet regions. Nat. Clim. Chang., 6, 508–513, doi:10.1038/nclimate2941.
  595. Gonzalez-Martin, C., N. Teigell-Perez, B. Valladares, and D.W. Griffin, 2014: The global dispersion of pathogenic microorganisms by dust storms and its relevance to agriculture. Adv. Agron., 127, 1–41, doi:10.1016/B978-0-12-800131-8.00001-7.
  596. Goudie, A.S., and N.J. Middleton, 2001: Saharan dust storms: Nature and consequences. Earth-Science Rev., 56, 179–204.
  597. Middleton, N., and U. Kang, 2017: Sand and dust storms: Impact mitigation. Sustainability, 9, 1053, doi:10.3390/su9061053.
  598. Wiesmeier, M., S. Munro, F. Barthold, M. Steffens, P. Schad, and 
I. Kögel-Knabner, 2015: Carbon storage capacity of semi-arid grassland soils and sequestration potentials in northern China. Glob. Chang. Biol., 21, 3836–3845, doi:10.1111/gcb.12957.
  599. Field, J.P. et al., 2010: The ecology of dust. Front. Ecol. Environ., 8, 423–430, doi:10.1890/090050.
  600. Stefanski, R., and M.V.K. Sivakumar, 2009: Impacts of sand and dust storms on agriculture and potential agricultural applications of a SDSWS. IOP Conf. Ser. Earth Environ. Sci., 7, 12016, doi:10.1088/1755-1307/7/1/012016.
  601. Middleton, N., and U. Kang, 2017: Sand and dust storms: Impact mitigation. Sustainability, 9, 1053, doi:10.3390/su9061053.
  602. Middleton, N.J., 2017: Desert dust hazards: A global review. Aeolian Res., 24, 53–63, doi:10.1016/J.AEOLIA.2016.12.001.
  603. Stefanski, R., and M.V.K. Sivakumar, 2009: Impacts of sand and dust storms on agriculture and potential agricultural applications of a SDSWS. IOP Conf. Ser. Earth Environ. Sci., 7, 12016, doi:10.1088/1755-1307/7/1/012016.
  604. Stefanski, R., and M.V.K. Sivakumar, 2009: Impacts of sand and dust storms on agriculture and potential agricultural applications of a SDSWS. IOP Conf. Ser. Earth Environ. Sci., 7, 12016, doi:10.1088/1755-1307/7/1/012016.
  605. Asem, S.O., and W.Y. Roy, 2010: Biodiversity and climate change in Kuwait. Int. J. Clim. Chang. Strateg. Manag., 2, 68–83, doi:10.1108/17568691011020265.
  606. Womack, A.M., B.J.M. Bohannan, and J.L. Green, 2010: Biodiversity and biogeography of the atmosphere. Philos. Trans. R. Soc. B Biol. Sci., 365, 3645–3653, doi:10.1098/rstb.2010.0283.
  607. Kellogg, C.A. et al., 2004: Characterization of aerosolized bacteria and fungi from desert dust events in Mali, West Africa. Aerobiologia, 20, 99–110, doi:10.1023/B:AERO.0000032947.88335.bb 99–110.
  608. Prospero, J.M., E. Blades, G. Mathison, and R. Naidu, 2005: Interhemispheric transport of viable fungi and bacteria from Africa to the Caribbean with soil dust. Aerobiologia (Bologna)., 21, 1–19, doi:10.1007/s10453-004-5872-7.
  609. Griffin, D.W., D.L. Westphal, and M.A. Gray, 2006: Airborne microorganisms in the African desert dust corridor over the mid-Atlantic ridge, Ocean Drilling Program, Leg 209. Aerobiologia (Bologna)., 22, 211–226, doi:10.1007/s10453-006-9033-z.
  610. Schlesinger, P., Y. Mamane, and I. Grishkan, 2006: Transport of microorganisms to Israel during Saharan dust events. Aerobiologia (Bologna)., 22, 259–273, doi:10.1007/s10453-006-9038-7.
  611. Griffin, D.W., 2007: Atmospheric movement of microorganisms in clouds of desert dust and implications for human health. Clin. Microbiol. Rev., 20, 459–477, doi:10.1128/CMR.00039-06.
  612. de Deckker, P. et al., 2008: Geochemical and microbiological fingerprinting of airborne dust that fell in Canberra, Australia, in October 2002. Geochemistry, Geophys. Geosystems, 9, doi:10.1029/2008GC002091.
  613. Jeon, E.M. et al., 2011: Impact of Asian dust events on airborne bacterial community assessed by molecular analyses. Atmos. Environ., 45, 4313–4321, doi:10.1016/j.atmosenv.2010.11.054.
  614. Abed, R.M.M., A. Ramette, V. Hübner, P. de Deckker, and D. de Beer, 2012: Microbial diversity of eolian dust sources from saline lake sediments and biological soil crusts in arid southern Australia. FEMS Microbiol. Ecol., 80, 294–304, doi:10.1111/j.1574-6941.2011.01289.x.
  615. Favet, J. et al., 2013: Microbial hitchhikers on intercontinental dust: Catching a lift in Chad. ISME J., 7, 850–867, doi:10.1038/ismej.2012.152.
  616. Woo, A.C. et al., 2013: Temporal variation in airborne microbial populations and microbially-derived allergens in a tropical urban landscape. Atmos. Environ., 74, 291–300, doi:10.1016/j.atmosenv.2013.03.047.
  617. Pointing, S.B., and J. Belnap, 2014: Disturbance to desert soil ecosystems contributes to dust-mediated impacts at regional scales. Biodivers. Conserv., 23, 1659–1667, doi:10.1007/s10531-014-0690-x.
  618. White, R., and J. Nackoney, 2003: Drylands, People, and Ecosystem Goods and Services: A Web-Based Geospatial Analysis. World Resource Institute, Washington, DC, USA, 40 pp.
  619. Martínez-Palacios, A., L.E. Eguiarte, and G.R. Furnier, 1999: Genetic diversity of the endangered endemic Agave victoriae-reginae (Agavaceae) in the Chihuahuan Desert. Am. J. Bot., 86, 1093–1098, doi:10.2307/2656971.
  620. Millennium Ecosystem Assessment, 2005: Ecosystems and Human Well-Being: Desertification Synthesis. World Resource Institute, Washington, DC, 
USA, 26 pp.
  621. Maestre, F.T. et al., 2012: Plant species richness and ecosystems multifunctionality in global drylands. Science, 335, 2014–2017, doi:10.1126/science.1215442.
  622. Gitay, H., A. Suárez, R.T. Watson, and D.J. Dokken, 2002: Climate change and biodiversity. IPCC Technical Paper V, Intergovernmental Panel on Climate Change, World Meteorological Organization, Geneva, Switzerland. 77 pp.
  623. Newbold, T. et al., 2015: Global effects of land use on local terrestrial biodiversity. Nature, 520, 45–50, doi:10.1038/nature14324.
  624. Wilting, H.C., A.M. Schipper, M. Bakkenes, J.R. Meijer, and M.A.J. Huijbregts, 2017: Quantifying biodiversity losses due to human consumption: A global-scale footprint analysis. Environ. Sci. Technol., 51, 3298–3306, doi:10.1021/acs.est.6b05296.
  625. Observatoire du Sahara et du Sahel, 2013: La Surveillance environnementale dans le circum-Sahara: Synthèse régionale Ecologie (Algérie, Burkina Faso – Kenya – Mali Niger – Sénégal – Tunisie) 2012, Observatoire du Sahara et du Sahel.
  626. Akhter, R., and M. Arshad, 2006: Arid rangelands in the Cholistan Desert (Pakistan). Artic. Sci., 17, 210–217.
  627. Khishigbayar, J. et al., 2015: Mongolian rangelands at a tipping point? Biomass and cover are stable but composition shifts and richness declines after 20 years of grazing and increasing temperatures. J. Arid Environ., 115, 100–112, doi:10.1016/j.jaridenv.2015.01.007.
  628. Ivanov, A.L. et al., 2018: National report global climate and soil cover of Russia: Assessment of risks and environmental and economic consequences of land degradation. Adaptive systems and technologies of environmental management (agriculture and forestry). In: Russian: NATSIONAL’N. Moscow, Russia, 357 pp.
  629. Vetter, M., C. Wirth, H. Bottcher, G. Churkina, E.-D. Schulze, T. Wutzler, and G. Weber, 2005: Partitioning direct and indirect human-induced effects on carbon sequestration of managed coniferous forests using model simulations and forest inventories. Glob. Chang. Biol., 11, 810–827, doi:10.1111/j.1365-2486.2005.00932.x.
  630. Vetter, S., 2005: Rangelands at equilibrium and non-equilibrium: Recent developments in the debate. J. Arid Environ., 62, 321–341, doi:10.1016/J.JARIDENV.2004.11.015.
  631. Hiernaux, P., and H.N. Le Houérou, 2006: The rangelands of the Sahel. Sécheresse 17(1) 17, pp. 51–71.
  632. Le Houérou, H.N., 1996: Climate change, drought and desertification. J. Arid Environ., 34, 133–185, doi:10.1006/jare.1996.0099.
  633. Hirche, A. et al., 2018: The Maghreb (North Africa) rangelands evolution for forty years: Regreening or degradation? In: Desertification: Past, Current and Future Trends [Squires, V.R. and A. Ariapour (eds.)]. Nova Science Publishers, New York, USA, pp. 73–106.
  634. Observatoire du Sahara et du Sahel, 2013: La Surveillance environnementale dans le circum-Sahara: Synthèse régionale Ecologie (Algérie, Burkina Faso – Kenya – Mali Niger – Sénégal – Tunisie) 2012, Observatoire du Sahara et du Sahel.
  635. Ivanov, A.L. et al., 2018: National report global climate and soil cover of Russia: Assessment of risks and environmental and economic consequences of land degradation. Adaptive systems and technologies of environmental management (agriculture and forestry). In: Russian: NATSIONAL’N. Moscow, Russia, 357 pp.
  636. Millennium Ecosystem Assessment, 2005: Ecosystems and Human Well-Being: Desertification Synthesis. World Resource Institute, Washington, DC, 
USA, 26 pp.
  637. Whitford, W., 2002: Ecology of Desert Systems. Academic Press, San Diego, USA, 343 pp.
  638. Maestre, F.T. et al., 2012: Plant species richness and ecosystems multifunctionality in global drylands. Science, 335, 2014–2017, doi:10.1126/science.1215442.
  639. Millennium Ecosystem Assessment, 2005: Ecosystems and Human Well-Being: Desertification Synthesis. World Resource Institute, Washington, DC, 
USA, 26 pp.
  640. Durant, S.M. et al., 2014: Fiddling in biodiversity hotspots while deserts burn? Collapse of the Sahara’s megafauna. Divers. Distrib., 20, 114–122, doi:10.1111/ddi.12157.
  641. Walther, B.A., 2016: A review of recent ecological changes in the Sahel, with particular reference to land use change, plants, birds and mammals. Afr. J. Ecol., 54, 268–280, doi:10.1111/aje.12350.
  642. Ceballos, G. et al., 2010: Rapid decline of a grassland system and its ecological and conservation implications. PLoS One, 5, e8562, doi:10.1371/journal.pone.0008562.
  643. Tang, Z., H. An, G. Zhu, and Z. Shangguan, 2018: Beta diversity diminishes in a chronosequence of desertification in a desert steppe. L. Degrad. Dev., 29, 543–550, doi:10.1002/ldr.2885.
  644. Newbold, T. et al., 2015: Global effects of land use on local terrestrial biodiversity. Nature, 520, 45–50, doi:10.1038/nature14324.
  645. Parsons, S.A., A. Kutt, E.P. Vanderduys, J.J. Perry, and L. Schwarzkopf, 2017: Exploring relationships between native vertebrate biodiversity and grazing land condition. Rangel. J., 39, 25–37, doi:10.1071/RJ16049.
  646. Whitford, W.G., 1997: Desertification and animal biodiversity in the desert grasslands of North America. J. Arid Environ., 37, 709–720, doi:10.1006/jare.1997.0313.
  647. Parsons, S.A., A. Kutt, E.P. Vanderduys, J.J. Perry, and L. Schwarzkopf, 2017: Exploring relationships between native vertebrate biodiversity and grazing land condition. Rangel. J., 39, 25–37, doi:10.1071/RJ16049.
  648. Schooley, R.L., B.T. Bestelmeyer, and A. Campanella, 2018: Shrub encroachment, productivity pulses, and core-transient dynamics of Chihuahuan Desert rodents. Ecosphere, 9, doi:10.1002/ecs2.2330.
  649. Hetem, R.S., S.K. Maloney, A. Fuller, and D. Mitchell, 2016: Heterothermy in large mammals: Inevitable or implemented? Biol. Rev., 91, 187–205, doi:10.1111/brv.12166.
  650. Albright, T.P. et al., 2017: Mapping evaporative water loss in desert passerines reveals an expanding threat of lethal dehydration. Proc. Natl. Acad. Sci., 114, 2283–2288, doi:10.1073/pnas.1613625114.
  651. Rey, B., A. Fuller, D. Mitchell, L.C.R. Meyer, and R.S. Hetem, 2017: Drought-induced starvation of aardvarks in the Kalahari: An indirect effect of climate change. Biol. Lett., 13, 20170301, doi:10.1098/rsbl.2017.0301.
  652. Walther, G.R., 2010: Community and ecosystem responses to recent climate change. Philos. Trans. R. Soc. B Biol. Sci., 365, 2019–2024, doi:10.1098/rstb.2010.0021.
  653. Zinabu, S., A. Kebede, B. Ferede, and J. Dugassa, 2018: Review on the relationship of climate change and prevalence of animal diseases. World Journal of Veterinary Science, 6, 6–18, doi:10.12970/2310-0796.2018.06.02.
  654. Fuller, A., D. Mitchell, S.K. Maloney, and R.S. Hetem, 2016: Towards a mechanistic understanding of the responses of large terrestrial mammals to heat and aridity associated with climate change. Clim. Chang. Responses, 3, 10, doi:10.1186/s40665-016-0024-1.
  655. UN, 2015: Outcomes and Policy-Oriented Recommendations from the UNCCD 3rd Scientific Conference. United Nations Convention to Combat Desertification, Twelfth Session, Committee on Science and Technology, Ankara, Turkey, 15 pp.
  656. Pradhan, P., L. Costa, D. Rybski, W. Lucht, and J.P. Kropp, 2017: A systematic study of sustainable development goal (SDG) interactions. Earth’s Futur., 5, 1169–1179, doi:10.1002/2017EF000632.
  657. Olsson, L., M. Opondo, P. Tschakert, A. Agrawal, S.H. Eriksen, S. Ma, L.N. Perch, and S.A. Zakieldeen, 2014: Livelihoods and Poverty. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White, (eds.)]. Cambridge University Press, 
New York, USA. pp. 793–832.
  658. Gerber, N., E. Nkonya, and J. von Braun, 2014: Land degradation, poverty and marginality. In: Marginality: Addressing the nexus of poverty, exclusion and ecology [Von Braun, J., and F.W. Gatzweiler (eds.)]. Springer Netherlands, Dordrecht, Netherlands, pp. 181–202.
  659. Vu, Q.M., Q.B. Le, E. Frossard, and P.L.G. Vlek, 2014: Socio-economic and biophysical determinants of land degradation in Vietnam: An integrated causal analysis at the national level. Land Use Policy, 36, 605–617, doi:10.1016/J.LANDUSEPOL.2013.10.012.
  660. Way, S.-A., 2016: Examining the links between poverty and land degradation: From blaming the poor toward recognising the rights of the poor. In: Governing Global Desertification: Linking Environmental Degradation, Poverty and Participation [Johnson, P.M. (ed.)]. Routledge, London, UK, pp. 47–62, doi:10.4324/9781315253916-13.
  661. Barbier, E.B., and J.P. Hochard, 2016: Does land degradation increase poverty in developing countries? PLoS One, 11, e0152973, doi:10.1371/journal.pone.0152973.
  662. Diao, X., and D.B. Sarpong, 2011: Poverty implications of agricultural land degradation in Ghana: An economy-wide, multimarket model assessment. African Dev. Rev., 23, 263–275, doi:10.1111/j.1467-8268.2011.00285.x.
  663. Kirui, O.K., 2016: Economics of Land Degradation, Sustainable Land Management and Poverty in Eastern Africa – The Extent, Drivers, Costs and Impacts. PhD Thesis, University of Bonn, Bonn, Germany, 165 pp.
  664. Jiang, Z., Y. Lian, and X. Qin, 2014: Rocky desertification in Southwest China: Impacts, causes, and restoration. Earth-Science Rev., 132, 1–12, doi:10.1016/J.EARSCIREV.2014.01.005.
  665. Ge, X., Y. Li, A.E. Luloff, K. Dong, and J. Xiao, 2015: Effect of agricultural economic growth on sandy desertification in Horqin Sandy Land. Ecol. Econ., 119, 53–63, doi:10.1016/J.ECOLECON.2015.08.006.
  666. Stern, D.I., 2017: The environmental Kuznets curve after 25 years. J. Bioeconomics, 19, 7–28, doi:10.1007/s10818-017-9243-1.
  667. FAO, 2018: The Impact of Disasters and Crises on Agriculture and Food Security 2017. Food and Agriculture Organization of the United Nations, Rome, Italy, 143 pp.
  668. FAO, 2018: The Impact of Disasters and Crises on Agriculture and Food Security 2017. Food and Agriculture Organization of the United Nations, Rome, Italy, 143 pp.
  669. FAO, 2018: The Impact of Disasters and Crises on Agriculture and Food Security 2017. Food and Agriculture Organization of the United Nations, Rome, Italy, 143 pp.
  670. Hochman, Z., D.L. Gobbett, and H. Horan, 2017: Climate trends account for stalled wheat yields in Australia since 1990. Glob. Chang. Biol., 23, 2071–2081, doi:10.1111/gcb.13604.
  671. Nelson, G.C. et al., 2010: Food Security, Farming, and Climate Change to 2050: Scenarios, Results, Policy Options. International Food Policy Research Institute, Washington, DC, USA. DOI: 10.2499/9780896291867, 131 pp.
  672. Zhao, C. et al., 2017: Temperature increase reduces global yields of major crops in four independent estimates. Proc. Natl. Acad. Sci. U.S.A., 114, 9326–9331, doi:10.1073/pnas.1701762114.
  673. Kirui, O.K., 2016: Economics of Land Degradation, Sustainable Land Management and Poverty in Eastern Africa – The Extent, Drivers, Costs and Impacts. PhD Thesis, University of Bonn, Bonn, Germany, 165 pp.
  674. Moussa, B., E. Nkonya, S. Meyer, E. Kato, T. Johnson, and J. Hawkins, 2016: Economics of land degradation and improvement in Niger. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 499–539.
  675. Mythili, G., and J. Goedecke, 2016: Economics of Land Degradation in India. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 431–469.
  676. Tun, K.K.K., R.P. Shrestha, and A. Datta, 2015: Assessment of land degradation and its impact on crop production in the dry zone of Myanmar. Int. J. Sustain. Dev. World Ecol., 22, 533–544, doi:10.1080/13504509.2015.1091046.
  677. Nkonya, E. et al., 2016a: Global cost of land degradation. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 117–165.
  678. Jish Prakash, P., G. Stenchikov, S. Kalenderski, S. Osipov, and H. Bangalath, 2015: The impact of dust storms on the Arabian Peninsula and the Red Sea. Atmos. Chem. Phys., 15, 199–222, doi:10.5194/acp-15-199-2015.
  679. Yu, K., P. D’Odorico, A. Bhattachan, G.S. Okin, and A.T. Evan, 2015: Dust-rainfall feedback in West African Sahel. Geophys. Res. Lett., 42, 7563–7571, doi:10.1002/2015GL065533.
  680. Gherboudj, I., S. Naseema Beegum, and H. Ghedira, 2017: Identifying natural dust source regions over the Middle-East and North-Africa: Estimation of dust emission potential. Earth-Science Rev., 165, 342–355, doi:10.1016/J.EARSCIREV.2016.12.010.
  681. Notaro, M., F. Alkolibi, E. Fadda, and F. Bakhrjy, 2013: Trajectory analysis of Saudi Arabian dust storms. J. Geophys. Res. Atmos., 118, 6028–6043, doi:10.1002/jgrd.50346.
  682. Yu, Y. et al., 2013: Assessing temporal and spatial variations in atmospheric dust over Saudi Arabia through satellite, radiometric, and station data. J. Geophys. Res. Atmos, 118, 253–266, doi:10.1002/2013JD020677.
  683. Alobaidi, M., M. Almazroui, A. Mashat, and P.D. Jones, 2017: Arabian Peninsula wet season dust storm distribution: Regionalization and trends analysis (1983-2013). Int. J. Climatol., 37, 1356–1373, doi:10.1002/joc.4782.
  684. Maghrabi, A., B. Alharbi, and N. Tapper, 2011: Impact of the March 2009 dust event in Saudi Arabia on aerosol optical properties, meteorological parameters, sky temperature and emissivity. Atmos. Environ., 45, 2164–2173, doi:10.1016/J.ATMOSENV.2011.01.071.
  685. Almazroui, M., M. Alobaidi, S. Saeed, A. Mashat, and M. Assiri, 2018: The possible impact of the circumglobal wave train on the wet season dust storm activity over the northern Arabian Peninsula. Clim. Dyn., 50, 2257–2268, doi:10.1007/s00382-017-3747-1.
  686. Rashki, A., D.G. Kaskaoutis, C.J. deW. Rautenbach, P.G. Eriksson, M. Qiang, and P. Gupta, 2012: Dust storms and their horizontal dust loading in the Sistan region, Iran. Aeolian Res., 5, 51–62, doi:10.1016/J.AEOLIA.2011.12.001.
  687. Türkeş, M.2017: Recent spatiotemporal variations of synoptic meteorological sand and dust storm events observed over the Middle East and surrounding regions. Proceedings of the 5th International Workshop on Sand and Dust Storms (SDS): Dust Sources and their Impacts in the Middle East, 23–25 October 2017. Istanbul, Turkey, pp. 45–59.
  688. Namdari, S., N. Karimi, A. Sorooshian, G. Mohammadi, and S. Sehatkashani, 2018: Impacts of climate and synoptic fluctuations on dust storm activity over the Middle East. Atmos. Environ., 173, 265–276, doi:10.1016/J.ATMOSENV.2017.11.016.
  689. Indoitu, R. et al., 2015: Dust emission and environmental changes in the dried bottom of the Aral Sea. Aeolian Res., 17, 101–115, doi:10.1016/J.AEOLIA.2015.02.004.
  690. Xi, X., and I.N. Sokolik, 2015: Seasonal dynamics of threshold friction velocity and dust emission in Central Asia. J. Geophys. Res. Atmos., 120, 1536–1564, doi:10.1002/2014JD022471.
  691. Díaz, J. et al., 2017: Saharan dust intrusions in Spain: Health impacts and associated synoptic conditions. Environ. Res., 156, 455–467, doi:10.1016/J.ENVRES.2017.03.047.
  692. Goudarzi, G. et al., 2017: Health risk assessment of exposure to the Middle-Eastern dust storms in the Iranian megacity of Kermanshah. Public Health, 148, 109–116, doi:10.1016/J.PUHE.2017.03.009.
  693. Goudie, A.S., 2014: Desert dust and human health disorders. Environ. Int., 63, 101–113, doi:10.1016/J.ENVINT.2013.10.011.
  694. Samoli, E., E. Kougea, P. Kassomenos, A. Analitis, and K. Katsouyanni, 2011: Does the presence of desert dust modify the effect of PM10 on mortality in Athens, Greece? Sci. Total Environ., 409, 2049–2054, doi:10.1016/J.SCITOTENV.2011.02.031.
  695. Goudie, A., and N. Middleton, 2006: Desert Dust in the Global System. Springer-Verlag, Berlin, Germany, 287 pp.
  696. Sprigg, W.A., 2016: Dust Storms, Human Health and a Global Early Warning System. Springer, Cham, Switzerland, 59–87 pp.
  697. Díaz, J. et al., 2017: Saharan dust intrusions in Spain: Health impacts and associated synoptic conditions. Environ. Res., 156, 455–467, doi:10.1016/J.ENVRES.2017.03.047.
  698. Goudarzi, G. et al., 2017: Health risk assessment of exposure to the Middle-Eastern dust storms in the Iranian megacity of Kermanshah. Public Health, 148, 109–116, doi:10.1016/J.PUHE.2017.03.009.
  699. Goudie, A.S., 2014: Desert dust and human health disorders. Environ. Int., 63, 101–113, doi:10.1016/J.ENVINT.2013.10.011.
  700. Samoli, E., E. Kougea, P. Kassomenos, A. Analitis, and K. Katsouyanni, 2011: Does the presence of desert dust modify the effect of PM10 on mortality in Athens, Greece? Sci. Total Environ., 409, 2049–2054, doi:10.1016/J.SCITOTENV.2011.02.031.
  701. Zhang, X. et al., 2016: A systematic review of global desert dust and associated human health effects. Atmosphere (Basel)., 7, 158, doi:10.3390/atmos7120158.
  702. Bennett, C.M., I.G. McKendry, S. Kelly, K. Denike, and T. Koch, 2006: Impact of the 1998 Gobi dust event on hospital admissions in the Lower Fraser Valley, British Columbia. Sci. Total Environ., 366, 918–925, doi:10.1016/j.scitotenv.2005.12.025.
  703. Díaz, J. et al., 2017: Saharan dust intrusions in Spain: Health impacts and associated synoptic conditions. Environ. Res., 156, 455–467, doi:10.1016/J.ENVRES.2017.03.047.
  704. Kashima, S., T. Yorifuji, S. Bae, Y. Honda, Y.-H. Lim, and Y.-C. Hong, 2016: Asian dust effect on cause-specific mortality in five cities across South Korea and Japan. Atmos. Environ., 128, 20–27, doi:10.1016/J.ATMOSENV.2015.12.063.
  705. Lee, H., Y. Honda, Y.-H. Lim, Y.L. Guo, M. Hashizume, and H. Kim, 2014: Effect of Asian dust storms on mortality in three Asian cities. Atmos. Environ., 89, 309–317, doi:10.1016/J.ATMOSENV.2014.02.048.
  706. Samoli, E., E. Kougea, P. Kassomenos, A. Analitis, and K. Katsouyanni, 2011: Does the presence of desert dust modify the effect of PM10 on mortality in Athens, Greece? Sci. Total Environ., 409, 2049–2054, doi:10.1016/J.SCITOTENV.2011.02.031.
  707. Zhang, X. et al., 2016: A systematic review of global desert dust and associated human health effects. Atmosphere (Basel)., 7, 158, doi:10.3390/atmos7120158.
  708. Goudie, A., 2013: The Human Impact on the Natural Environment: Past, Present and Future. Seventh edition, Wiley-Blackwell, West Sussex, UK, 422 pp.
  709. Giannadaki, D., A. Pozzer, and J. Lelieveld, 2014: Modeled global effects of airborne desert dust on air quality and premature mortality. Atmos. Chem. Phys., 14, 957–968, doi:10.5194/acp-14-957-2014.
  710. Giannadaki, D., A. Pozzer, and J. Lelieveld, 2014: Modeled global effects of airborne desert dust on air quality and premature mortality. Atmos. Chem. Phys., 14, 957–968, doi:10.5194/acp-14-957-2014.
  711. Kim, S.E. et al., 2017: Seasonal analysis of the short-term effects of air pollution on daily mortality in Northeast Asia. Sci. Total Environ., 576, 850–857, doi:10.1016/J.SCITOTENV.2016.10.036.
  712. Hopke, P.K. et al., 2018: Spatial and temporal trends of short-term health impacts of PM 2.5 in Iranian cities; A modelling approach (2013–2016). Aerosol Air Qual. Res., 18, 497–504, doi:10.4209/aaqr.2017.09.0325.
  713. Agier, L., A. Deroubaix, N. Martiny, P. Yaka, A. Djibo, and H. Broutin, 2012: Seasonality of meningitis in Africa and climate forcing: Aerosols stand out. J.R. Soc. Interface, 10, 20120814, doi:10.1098/rsif.2012.0814.
  714. Molesworth, A.M., L.E. Cuevas, S.J. Connor, A.P. Morse, and M.C. Thomson, 2003: Environmental risk and meningitis epidemics in Africa. Emerg. Infect. Dis., 9, 1287–1293, doi:10.3201/eid0910.030182.
  715. Bose, P., India’s drylands agroforestry: A ten-year analysis of gender and social diversity, tenure and climate variability. Int. For. Rev., 17, doi:10.1505/146554815816086435.
  716. Broeckhoven, N., and A. Cliquet, 2015: Gender and ecological restoration: Time to connect the dots. Restor. Ecol., 23, 729–736, doi:10.1111/rec.12270.
  717. Kaijser, A., and A. Kronsell, 2014: Climate change through the lens of intersectionality. Env. Polit., 23, 417–433, doi:10.1080/09644016.2013.835203.
  718. Kiptot, E., S. Franzel, and A. Degrande, 2014: Gender, agroforestry and food security in Africa. Curr. Opin. Environ. Sustain., 6, 104–109, doi:10.1016/J.COSUST.2013.10.019.
  719. Villamor, G.B., and M. van Noordwijk, 2016: Gender specific land-use decisions and implications for ecosystem services in semi-matrilineal Sumatra. Glob. Environ. Chang., 39, 69–80, doi:10.1016/J.GLOENVCHA.2016.04.007.
  720. Arora-Jonsson, S., 2011: Virtue and vulnerability: Discourses on women, gender and climate change. Glob. Environ. Chang., 21, 744–751, doi:10.1016/J.GLOENVCHA.2011.01.005.
  721. Gurung, J.D., S. Mwanundu, A. Lubbock, M. Hartl, I. and Firmian, and IFAD, 2006: Gender and Desertification: Expanding Roles for Women to Restore Dryland Areas. International Fund for Agricultural Development (IFAD), Rome, Italy, 27 pp.
  722. Ahmed, A., E.T. Lawson, A. Mensah, C. Gordon, and J. Padgham, 2016: Adaptation to climate change or non-climatic stressors in semi-arid regions? Evidence of gender differentiation in three agrarian districts of Ghana. Environ. Dev., 20, 45–58, doi:10.1016/J.ENVDEV.2016.08.002.
  723. Kaijser, A., and A. Kronsell, 2014: Climate change through the lens of intersectionality. Env. Polit., 23, 417–433, doi:10.1080/09644016.2013.835203.
  724. Villamor, G.B., M. van Noordwijk, U. Djanibekov, M.E. Chiong-Javier, and D. Catacutan, 2014: Gender differences in land-use decisions: Shaping multifunctional landscapes? Curr. Opin. Environ. Sustain., 6, 128–133, doi:10.1016/J.COSUST.2013.11.015.
  725. Sijapati Basnett, B., M. Elias, M. Ihalainen, and A.M. Paez Valencia, 2017: Gender Matters in Forest Landscape Restoration: A framework for Design and Evaluation. Center for International Forestry Research, Bogor, Indonesia, 12 pp.
  726. Catacutan, D.C., and G.B. Villamor, 2016: Gender roles and land use preferences – Implications to landscape restoration in Southeast Asia. In: Land Restoration: Reclaiming Landscapes for a Sustainable Future. [Chabay, I., M. Frick and J. Helgeson (eds.)]. Elsevier, Oxford, UK, pp. 431–440.
  727. De la Paix, M.J., L. Lanhai, C. Xi, S. Ahmed, and A. Varenyam, 2011: Soil degradation and altered flood risk as a consequence of deforestation. L. Degrad. Dev., 24, 478-485, doi:10.1002/ldr.1147, pp. 478–485.
  728. le Roux, B., M. van der Laan, T. Vahrmeijer, K.L. Bristow, and J.G. Annandale, 2017: Establishing and testing a catchment water footprint framework to inform sustainable irrigation water use for an aquifer under stress. Sci. Total Environ., 599–600, 1119–1129, doi:10.1016/J.SCITOTENV.2017.04.170.
  729. Cano, A. et al., 2018: Current knowledge and future research directions to link soil health and water conservation in the Ogallala Aquifer region. Geoderma, 328, 109–118, doi:10.1016/J.GEODERMA.2018.04.027.
  730. Cook, B.I., R.L. Miller, and R. Seager, 2009: Amplification of the North American ‘Dust Bowl’ drought through human-induced land degradation. Proc. Natl. Acad. Sci. U.S.A., 106, 4997–5001, doi:10.1073/pnas.0810200106.
  731. Greene, R., W. Timms, P. Rengasamy, M. Arshad, and R. Cresswell, 2016: Soil and aquifer salinization: Toward an integrated approach for salinity management of groundwater. In: Integrated Groundwater Management [Jakeman, A.J., O. Barreteau, R.J. Hunt, J. Rinaudo and A. Ross (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 377–412.
  732. Wichelns, D., and M. Qadir, 2015: Achieving sustainable irrigation requires effective management of salts, soil salinity, and shallow groundwater. Agric. Water Manag., 157, 31–38, doi:10.1016/J.AGWAT.2014.08.016.
  733. Liu, Y., J. Liu, and Y. Zhou, 2017: Spatio-temporal patterns of rural poverty in China and targeted poverty alleviation strategies. J. Rural Stud., 52, 66–75, doi:10.1016/J.JRURSTUD.2017.04.002.
  734. Liu, J., Q. Liu, and H. Yang, 2016c: Assessing water scarcity by simultaneously considering environmental flow requirements, water quantity, and water quality. Ecol. Indic., 60, 434–441, doi:10.1016/J.ECOLIND.2015.07.019.
  735. IPCC, 2013: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, USA.
  736. Zheng, J., Y. Yu, X. Zhang, and Z. Hao, 2018: Variation of extreme drought and flood in North China revealed by document-based seasonal precipitation reconstruction for the past 300 years. Clim. Past, 14, 1135–1145, doi:10.5194/cp-14-1135-2018.
  737. Pedrero, F., I. Kalavrouziotis, J.J. Alarcón, P. Koukoulakis, and T. Asano, 2010: Use of treated municipal wastewater in irrigated agriculture – Review of some practices in Spain and Greece. Agric. Water Manag., 97, 1233–1241, doi:10.1016/J.AGWAT.2010.03.003.
  738. Tal, A., 2016: Rethinking the sustainability of Israel’s irrigation practices in the drylands. Water Res., 90, 387–394, doi:10.1016/J.WATRES.2015.12.016.
  739. Singh, K.P., D. Mohan, S. Sinha, and R. Dalwani, 2004: Impact assessment of treated/untreated wastewater toxicants discharged by sewage treatment plants on health, agricultural, and environmental quality in the wastewater disposal area. Chemosphere, 55, 227–255, doi:10.1016/J.CHEMOSPHERE.2003.10.050.
  740. Qishlaqi, A., F. Moore, and G. Forghani, 2008: Impact of untreated wastewater irrigation on soils and crops in Shiraz suburban area, SW Iran. Environ. Monit. Assess., 141, 257–273, doi:10.1007/s10661-007-9893-x.
  741. Hanjra, M.A., J. Blackwell, G. Carr, F. Zhang, and T.M. Jackson, 2012: Wastewater irrigation and environmental health: Implications for water governance and public policy. Int. J. Hyg. Environ. Health, 215, 255–269, doi:10.1016/J.IJHEH.2011.10.003.
  742. Faour-Klingbeil, D., and E.C.D. Todd, 2018: The impact of climate change on raw and untreated wastewater use for agriculture, especially in arid regions: A review. Foodborne Pathog. Dis., 15, 61–72, doi:10.1089/fpd.2017.2389.
  743. Hanjra, M.A., J. Blackwell, G. Carr, F. Zhang, and T.M. Jackson, 2012: Wastewater irrigation and environmental health: Implications for water governance and public policy. Int. J. Hyg. Environ. Health, 215, 255–269, doi:10.1016/J.IJHEH.2011.10.003.
  744. Zidane, I.F., K.M. Saqr, G. Swadener, X. Ma, and M.F. Shehadeh, 2016: On the role of surface roughness in the aerodynamic performance and energy conversion of horizontal wind turbine blades: A review. Int. J. Energy Res., 40, 2054–2077, doi:10.1002/er.3580.
  745. Costa, S.C.S., A.S.A.C. Diniz, and L.L. Kazmerski, 2016: Dust and soiling issues and impacts relating to solar energy systems: Literature review update for 2012–2015. Renew. Sustain. Energy Rev., 63, 33–61, doi:10.1016/J.RSER.2016.04.059.
  746. Lopez-Garcia, J., A. Pozza, and T. Sample, 2016: Long-term soiling of silicon PV modules in a moderate subtropical climate. Sol. Energy, 130, 174–183, doi:10.1016/J.SOLENER.2016.02.025.
  747. Maliszewski, P.J., E.K. Larson, and C. Perrings, 2012: Environmental determinants of unscheduled residential outages in the electrical power distribution of Phoenix, Arizona. Reliab. Eng. Syst. Saf., 99, 161–171, doi:10.1016/J.RESS.2011.10.011.
  748. Mani, M., and R. Pillai, 2010: Impact of dust on solar photovoltaic (PV) performance: Research status, challenges and recommendations. Renew. Sustain. Energy Rev., 14, 3124–3131, doi:10.1016/J.RSER.2010.07.065.
  749. Mejia, F.A., and J. Kleissl, 2013: Soiling losses for solar photovoltaic systems in California. Sol. Energy, 95, 357–363, doi:10.1016/J.SOLENER.2013.06.028.
  750. Mejia, F., J. Kleissl, and J.L. Bosch, 2014: The effect of dust on solar photovoltaic systems. Energy Procedia, 49, 2370–2376, doi:10.1016/J.EGYPRO.2014.03.251.
  751. Middleton, N.J., 2017: Desert dust hazards: A global review. Aeolian Res., 24, 53–63, doi:10.1016/J.AEOLIA.2016.12.001.
  752. Sarver, T., A. Al-Qaraghuli, and L.L. Kazmerski, 2013: A comprehensive review of the impact of dust on the use of solar energy: History, investigations, results, literature, and mitigation approaches. Renew. Sustain. Energy Rev., 22, 698–733, doi:10.1016/J.RSER.2012.12.065.
  753. Kaufman, Y.J., D. Tanré, and O. Boucher, 2002: A satellite view of aerosols in the climate system. Nature, 419, 215–223, doi:10.1038/nature01091.
  754. Kok, J.F., D.S. Ward, N.M. Mahowald, and A.T. Evan, 2018: Global and regional importance of the direct dust-climate feedback. Nat. Commun., 9, 241, doi:10.1038/s41467-017-02620-y.
  755. Ghazi, S., A. Sayigh, and K. Ip, 2014: Dust effect on flat surfaces – A review paper. Renew. Sustain. Energy Rev., 33, 742–751, doi:10.1016/J.RSER.2014.02.016.
  756. Saidan, M., A.G. Albaali, E. Alasis, and J.K. Kaldellis, 2016: Experimental study on the effect of dust deposition on solar photovoltaic panels in desert environment. Renew. Energy, 92, 499–505, doi:10.1016/J.RENENE.2016.02.031.
  757. Fuentealba, E. et al., 2015: Photovoltaic performance and LCoE comparison at the coastal zone of the Atacama Desert, Chile. Energy Convers. Manag., 95, 181–186, doi:10.1016/J.ENCONMAN.2015.02.036.
  758. Middleton, N.J., 2017: Desert dust hazards: A global review. Aeolian Res., 24, 53–63, doi:10.1016/J.AEOLIA.2016.12.001.
  759. Costa, S.C.S., A.S.A.C. Diniz, and L.L. Kazmerski, 2016: Dust and soiling issues and impacts relating to solar energy systems: Literature review update for 2012–2015. Renew. Sustain. Energy Rev., 63, 33–61, doi:10.1016/J.RSER.2016.04.059.
  760. Costa, C.S.C., Suellen, A.S.A.C. Diniz, and L.L. Kazmerski, 2018: Solar energy dust and soiling R&D progress: Literature review update for 2016. Renew. Sustain. Energy Rev., 82, 2504–2536, doi:10.1016/J.RSER.2017.09.015.
  761. Gholami, A., A.A. Alemrajabi, and A. Saboonchi, 2017: Experimental study of self-cleaning property of titanium dioxide and nanospray coatings in solar applications. Sol. Energy, 157, 559–565, doi:10.1016/J.SOLENER.2017.08.075.
  762. Maghrabi, A., B. Alharbi, and N. Tapper, 2011: Impact of the March 2009 dust event in Saudi Arabia on aerosol optical properties, meteorological parameters, sky temperature and emissivity. Atmos. Environ., 45, 2164–2173, doi:10.1016/J.ATMOSENV.2011.01.071.
  763. Bruno, L., M. Horvat, and L. Raffaele, 2018: Windblown sand along railway infrastructures: A review of challenges and mitigation measures. J. Wind Eng. Ind. Aerodyn., 177, 340–365, doi:10.1016/J.JWEIA.2018.04.021.
  764. Bruno, L., M. Horvat, and L. Raffaele, 2018: Windblown sand along railway infrastructures: A review of challenges and mitigation measures. J. Wind Eng. Ind. Aerodyn., 177, 340–365, doi:10.1016/J.JWEIA.2018.04.021.
  765. Cheng, J., and C. Xue, 2014: The sand-damage-prevention engineering system for the railway in the desert region of the Qinghai-Tibet plateau. J. Wind Eng. Ind. Aerodyn., 125, 30–37, doi:10.1016/J.JWEIA.2013.11.016.
  766. Bruno, L., M. Horvat, and L. Raffaele, 2018: Windblown sand along railway infrastructures: A review of challenges and mitigation measures. J. Wind Eng. Ind. Aerodyn., 177, 340–365, doi:10.1016/J.JWEIA.2018.04.021.
  767. Zhang, K., J. Qu, K. Liao, Q. Niu, and Q. Han, 2010: Damage by wind-blown sand and its control along Qinghai-Tibet railway in China. Aeolian Res., 1, 143–146, doi:10.1016/J.AEOLIA.2009.10.001.
  768. Herrero, S.T., 2006: Desertification and environmental security: The case of conflicts between farmers and herders in the arid environments of the Sahel. In: Desertification in the Mediterranean Region. A Security Issue [Kepner W.G., J.L. Rubio, D.A. Mouat and F. Pedrazzini (eds.)]. NATO Security Series, Vol. 3, Springer Netherlands, Dordrecht, Netherlands, pp. 109–132.
  769. von Uexkull, N., M. Croicu, H. Fjelde, and H. Buhaug, 2016: Civil conflict sensitivity to growing-season drought. Proc. Natl. Acad. Sci. U.S.A., 113, 12391–12396, doi:10.1073/pnas.1607542113.
  770. Theisen, O.M., 2017: Climate change and violence: Insights from political science. Curr. Clim. Chang. Reports, 3, 210–221, doi:10.1007/s40641-017-0079-5.
  771. Olsson, L., 2017: Climate migration and conflicts. In: Climate change, Migration and Human Rights [Manou, D., A. Baldwin, D. Cubie, A. Mihr and T. Thorp (eds.)]. Routledge, London, UK, pp. 116–128.
  772. Wischnath, G., and H. Buhaug, 2014: Rice or riots: On food production and conflict severity across India. Polit. Geogr., 43, 6–15, doi:10.1016/J.POLGEO.2014.07.004.
  773. Hsiang, S.M., M. Burke, and E. Miguel, 2013: Quantifying the influence of climate on human conflict. Science, 341, doi:10.1126/science.1235367.
  774. Hsiang, S.M., M. Burke, and E. Miguel, 2013: Quantifying the influence of climate on human conflict. Science, 341, doi:10.1126/science.1235367.
  775. Buhaug, H. et al., 2014: One effect to rule them all? A comment on climate and conflict. Clim. Change, 127, 391–397, doi:10.1007/s10584-014-1266-1.
  776. Hsiang, S.M., M. Burke, and E. Miguel, 2013: Quantifying the influence of climate on human conflict. Science, 341, doi:10.1126/science.1235367.
  777. Almer, C., J. Laurent-Lucchetti, and M. Oechslin, 2017: Water scarcity and rioting: Disaggregated evidence from Sub-Saharan Africa. J. Environ. Econ. Manage., 86, 193–209, doi:10.1016/J.JEEM.2017.06.002.
  778. Owain, E.L., and M.A. Maslin, 2018: Assessing the relative contribution of economic, political and environmental factors on past conflict and the displacement of people in East Africa. Palgrave Commun., 4, 47, doi:10.1057/s41599-018-0096-6.
  779. Benjaminsen, T.A., 2016: Does climate change lead to conflicts in the Sahel? In: The End of Desertification? Disputing Environmental Change in the Drylands [Behnke R. and M. Mortimore, (eds.)]. Springer, Berlin, Heidelberg, Germany, pp. 99–116.
  780. Benjaminsen, T.A., K. Alinon, H. Buhaug, and J.T. Buseth, 2012: Does climate change drive land-use conflicts in the Sahel? J. Peace Res., 49, 97–111, doi:10.1177/0022343311427343.
  781. Verhoeven, H., 2011: Climate change, conflict and development in Sudan: Global neo-malthusian narratives and local power struggles. Dev. Change, 42, 679–707, doi:10.1111/j.1467-7660.2011.01707.x.
  782. de Châtel, F., 2014: The role of drought and climate change in the Syrian uprising: Untangling the triggers of the revolution. Middle East. Stud., 50, 521–535, doi:10.1080/00263206.2013.850076.
  783. Adams, C., T. Ide, J. Barnett, and A. Detges, 2018: Sampling bias in climate–conflict research. Nat. Clim. Chang., 8, 200–203, doi:10.1038/s41558-018-0068-2.
  784. World Bank, 2018: Groundswell: Preparing for Internal Climate Migration. World Bank Group, Washington DC, USA, 256 pp.
  785. Missirian, A., and W. Schlenker, 2017: Asylum applications respond to temperature fluctuations. Science, 358, 1610–1614, doi:10.1126/science.aao0432.
  786. Hunter, L.M., J.K. Luna, and R.M. Norton, 2015: Environmental dimensions of migration. Annu. Rev. Sociol., 41, 377–397, doi:10.1146/annurev-soc-073014-112223.
  787. McLeman, R.A., 2011: Settlement abandonment in the context of global environmental change. Glob. Environ. Chang., 21, S108–S120, doi:10.1016/J.GLOENVCHA.2011.08.004.
  788. De Sherbinin, A., and L. Bai, 2018: Geospatial modeling and mapping. In: Routledge Handbook of Environmental Displacement and Migration. Routledge, 85–91, doi:10.4324/9781315638843-6.
  789. Gemenne, F., 2011: Why the numbers don’t add up: A review of estimates and predictions of people displaced by environmental changes. Glob. Environ. Chang., 21, S41–S49, doi:10.1016/J.GLOENVCHA.2011.09.005.
  790. McLeman, R., 2013: Developments in modelling of climate change-related migration. Clim. Change, 117, 599–611, doi:10.1007/s10584-012-0578-2.
  791. Hartmann, B., 2010: Rethinking climate refugees and climate conflict: Rhetoric, reality and the politics of policy discourse. J. Int. Dev., 22, 233–246, doi:10.1002/jid.1676.
  792. Kniveton, D., C. Smith, and S. Wood, 2011: Agent-based model simulations of future changes in migration flows for Burkina Faso. Glob. Environ. Chang., 21, S34–S40, doi:10.1016/J.GLOENVCHA.2011.09.006.
  793. Piguet, E., 2010: Linking climate change, environmental degradation, and migration: A methodological overview. Wiley Interdiscip. Rev. Clim. Chang., 1, 517–524, doi:10.1002/wcc.54.
  794. Adger, W.N., J.M. Pulhin, J. Barnett, G.D. Dabelko, G.K. Hovelsrud, M. Levy, 
Ú.O. Spring, and C.H. Vogel, 2014: Human security. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, 
M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, 
A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White (eds.)]. Cambridge University Press, New York, USA, pp. 755–791.
  795. London Government Office for Science, 2011: Migration and Global Environmental Change: Future Challenges and Opportunities. Final Project Report, The Government Office for Science, London, UK, 235 pp.
  796. Melde, S., Laczko, F., and Gemenne, F., 2017: Making Mobility Work for Adaptation to Environmental Changes: Results from the MECLEP Global Research. International Organization for Migration, Geneva, Switzerland, 122 pp.
  797. Gray, C., and V. Mueller, 2012a: Drought and population mobility in rural Ethiopia. World Dev., 40, 134–145, doi:10.1016/J.WORLDDEV.2011.05.023.
  798. Gray, C.L., and V. Mueller, 2012b: Natural disasters and population mobility in Bangladesh. Proc. Natl. Acad. Sci. U.S.A., 109, 6000–6005, doi:10.1073/pnas.1115944109.
  799. Mueller, V., C. Gray, and K. Kosec, 2014: Heat stress increases long-term human migration in rural Pakistan. Nat. Clim. Chang., 4, 182–185, doi:10.1038/nclimate2103.
  800. Halliday, T., 2006: Migration, risk, and liquidity constraints in El Salvador. Econ. Dev. Cult. Change, 54, 893–925, doi:10.1086/503584.
  801. Gray, C., and V. Mueller, 2012a: Drought and population mobility in rural Ethiopia. World Dev., 40, 134–145, doi:10.1016/J.WORLDDEV.2011.05.023.
  802. Missirian, A., and W. Schlenker, 2017: Asylum applications respond to temperature fluctuations. Science, 358, 1610–1614, doi:10.1126/science.aao0432.
  803. Mueller, V., C. Gray, and K. Kosec, 2014: Heat stress increases long-term human migration in rural Pakistan. Nat. Clim. Chang., 4, 182–185, doi:10.1038/nclimate2103.
  804. De Sherbinin, A. et al., 2012: Migration and risk: Net migration in marginal ecosystems and hazardous areas. Environ. Res. Lett., 7, 45602, doi:10.1088/1748-9326/7/4/045602.
  805. Ezra, M., and G.-E. Kiros, 2001: Rural out-migration in the drought prone areas of Ethiopia: A multilevel analysis. Int. Migr. Rev., 35, 749–771, doi:10.1111/j.1747-7379.2001.tb00039.x.
  806. Findlay, A.M., 2011: Migrant destinations in an era of environmental change. Glob. Environ. Chang., 21, S50–S58, doi:10.1016/J.GLOENVCHA.2011.09.004.
  807. Gray, C., and V. Mueller, 2012a: Drought and population mobility in rural Ethiopia. World Dev., 40, 134–145, doi:10.1016/J.WORLDDEV.2011.05.023.
  808. Henry, S., B. Schoumaker, and C. Beauchemin, 2004: The impact of rainfall on the first out-migration: A multi-level event-history analysis in Burkina Faso. Popul. Environ., 25, 423–460, doi:10.1023/B:POEN.0000036928.17696.e8.
  809. Feng, S., Krueger, A.B., and Oppenheimer, M., 2010: Linkages among climate change, crop yields and Mexico–US cross-border migration. Natl. Acad Sci., 107, 14257–14262, doi:10.1073/pnas.1002632107.
  810. Nawrotzki, R.J., and J. DeWaard, 2016: Climate shocks and the timing of migration from Mexico. Popul. Environ., 38, 72–100, doi:10.1007/s11111-016-0255-x.
  811. Nawrotzki, R.J., F. Riosmena, L.M. Hunter, and D.M. Runfola, 2015: Undocumented migration in response to climate change. Int. J. Popul. Stud., 1, 60–74, doi:10.18063/IJPS.2015.01.004.
  812. Nawrotzki, R.J., and J. DeWaard, 2016: Climate shocks and the timing of migration from Mexico. Popul. Environ., 38, 72–100, doi:10.1007/s11111-016-0255-x.
  813. Gray, C., and V. Mueller, 2012a: Drought and population mobility in rural Ethiopia. World Dev., 40, 134–145, doi:10.1016/J.WORLDDEV.2011.05.023.
  814. Gray, C.L., 2011: Soil quality and human migration in Kenya and Uganda. Glob. Environ. Chang., 21, 421–430, doi:10.1016/J.GLOENVCHA.2011.02.004.
  815. Massey, D.S., W.G. Axinn, and D.J. Ghimire, 2010: Environmental change and out-migration: Evidence from Nepal. Popul. Environ., 32, 109–136, doi:10.1007/s11111-010-0119-8.
  816. Gray, C.L., 2011: Soil quality and human migration in Kenya and Uganda. Glob. Environ. Chang., 21, 421–430, doi:10.1016/J.GLOENVCHA.2011.02.004.
  817. Nawrotzki, R.J., and M. Bakhtsiyarava, 2017: International climate migration: Evidence for the climate inhibitor mechanism and the agricultural pathway. Popul. Space Place, 23, e2033, doi:10.1002/psp.2033.
  818. Barbier, E.B., and J.P. Hochard, 2018: Land degradation and poverty. Nat. Sustain., 1, 623–631, doi:10.1038/s41893-018-0155-4.
  819. Koubi, V., G. Spilker, L. Schaffer, and T. Bernauer, 2016: Environmental stressors and migration: Evidence from Vietnam. World Dev., 79, 197–210, doi:10.1016/J.WORLDDEV.2015.11.016.
  820. Kubik, Z., and M. Maurel, 2016: Weather shocks, agricultural production and migration: Evidence from Tanzania. J. Dev. Stud., 52, 665–680, doi:10.1080/00220388.2015.1107049.
  821. McKenzie, D., and D. Yang, 2015: Evidence on policies to increase the development impacts of international migration. World Bank Res. Obs., 30, 155–192, doi:10.1093/wbro/lkv001.
  822. Hugo, G., 2008: Migration, Development and Environment. IOM Migration Research Series, 35, International Organization for Migration, ISSN 1607-338X, Geneva, Switzerland. 63 pp.
  823. IPBES, 2018: Assessment Report on Land Degradation and Restoration [Montanarella, L., R. Scholes., and A. Brainich. (eds.)]. Secretariate of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, Bonn, Germany, 686 pp.
  824. McLeman, R., 2017: Migration and Land Degradation: Recent Experience and Future Trends. Global Land Outlook Working Paper, UNCCD: GLO, Bonn, Germany, 44 pp.
  825. Rass, N., 2006: Policies and Strategies to Address the Vulnerability of Pastoralists in Sub-Saharan Africa. PPLPI Working Paper No. 37, Food and Agriculture Organization of the United Nations, Rome, Italy, 22 pp.
  826. Dong, S., 2016: Overview: Pastoralism in the World. In: Building Resilience of Human-Natural Systems of Pastoralism in the Developing World [Dong, S., K.A.S. Kassam, J. Tourrand and Boone, R.B. (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 1–37.
  827. Gomes, N., 2006: Access to Water, Pastoral Resource Management and Pastoralists’ Livelihoods: Lessons Learned from Water Development in Selected Areas of Eastern Africa (Kenya, Ethiopia, Somalia). LSP Working Paper 26, Food and Agriculture Organization of the United Nations, Rome, Italy, 47 pp.
  828. Hein, L., and N. De Ridder, 2006: Desertification in the Sahel: A reinterpretation. Glob. Chang. Biol., 12, 751–758, doi:10.1111/j.1365-2486.2006.01135.x.
  829. Molua, E.L., and C.M. Lambi, 2007: The Economic Impact of Climate Change on Agriculture in Cameroon, Volume 1 of 1. Policy Research Working Paper, The World Bank, Washington, DC, doi:10.1596/1813-9450-4364.
  830. Kassahun, A., H.A. Snyman, and G.N. Smit, 2008: Impact of rangeland degradation on the pastoral production systems, livelihoods and perceptions of the Somali pastoralists in Eastern Ethiopia. J. Arid Environ., 72, 1265–1281, doi:10.1016/J.JARIDENV.2008.01.002.
  831. Vetter, S., 2005: Rangelands at equilibrium and non-equilibrium: Recent developments in the debate. J. Arid Environ., 62, 321–341, doi:10.1016/J.JARIDENV.2004.11.015.
  832. Hopkins, A., and A. Del Prado, 2007: Implications of climate change for grassland in Europe: Impacts, adaptations and mitigation options: A review. Grass Forage Sci., 62, 118–126, doi:10.1111/j.1365-2494.2007.00575.x.
  833. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  834. Dibari, C., M. Bindi, M. Moriondo, N. Staglianò, S. Targetti, and G. Argenti, 2016: Spatial data integration for the environmental characterisation of pasture macrotypes in the Italian Alps. Grass Forage Sci., 71, 219–234, doi:10.1111/gfs.12168.
  835. López-i-Gelats, F., E.D.G. Fraser, J.F. Morton, and M.G. Rivera-Ferre, 2016: What drives the vulnerability of pastoralists to global environmental change? A qualitative meta-analysis. Glob. Environ. Chang., 39, 258–274, doi:10.1016/J.GLOENVCHA.2016.05.011.
  836. Giannini, A., M. Biasutti, I.M. Held, and A. Sobel, 2008: A global perspective on African climate. Clim. Change, 90, 359–383, doi:10.1007/s10584-008-9396-y.
  837. IPCC, 2007: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, United Kingdom, 976 pp.
  838. Du, F., 2012: Ecological resettlement of Tibetan herders in the Sanjiangyuan: A case study in Madoi County of Qinghai. Nomad. People., 16, 116–133, doi:10.3167/np.2012.160109.
  839. Gebeye, B.A., 2016: Unsustain the sustainable: An evaluation of the legal and policy interventions for pastoral development in Ethiopia. Pastoralism, 6:2, doi:10.1186/s13570-016-0049-x.
  840. Liao, C., C. Barrett, and K.-A. Kassam, 2015: Does diversification improve livelihoods? Pastoral households in Xinjiang, China. Dev. Change, 46, 1302–1330, doi:10.1111/dech.12201.
  841. Schilling, J., Akuno, M., Scheffran, J., and Weinzierl, T., 2014: On raids and relations: Climate change, pastoral conflict and adaptation in north-western Kenya. In: Climate Change and Conflict: Where to for Conflict-Sensitive Adaptation to Climate Change in Africa? [Bronkhorst, S. and U. Bob (eds.)]. Human Research Council, Durban, South Africa. Pages 241–265.
  842. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  843. Otuoma, J., J. Kinyamario, W. Ekaya, M. Kshatriya, and M. Nyabenge, 2009: Effects of human-livestock-wildlife interactions on habitat in an Eastern Kenya rangeland. Afr. J. Ecol., 47, 567–573, doi:10.1111/j.1365-2028.2008.01009.x.
  844. Behnke, R., and C. Kerven, 2013: Counting the costs: Replacing pastoralism with irrigated agriculture in the Awash Valley, north-eastern Ethiopia. In: Pastoralism and Development in Africa: Dynamic Change at the Margins [Catley, A., J. Lind, and I. Scoones (eds.)]. Routledge, London, UK, pp. 328.
  845. Tamou, C., R. Ripoll-Bosch, I.J.M.M. de Boer, and S.J. Oosting, 2018: Pastoralists in a changing environment: The competition for grazing land in and around the W Biosphere Reserve, Benin Republic. Ambio, 47, 340–354, doi:10.1007/s13280-017-0942-6.
  846. Dodd, J., 1994: Desertification and degradation of Africa’s rangelands. Rangelands, 16(5), 180–183.
  847. Davies, J. et al., 2016: Improving Governance of Pastoral Lands: Implementing the Voluntary Guidelines on the Responsible Governance of Tenure Of Land, Fisheries and Forests in the Context of National Food Security. Governance of Tenure Technical Guide No. 6, Food and Agriculture Organization of the United Nations, Rome, 152 pp.
  848. Haan, C. De, E. Dubern, B. Garancher, and C. Quintero, 2016: Pastoralism Development in the Sahel: A Road to Stability? World Bank Group, Washington DC, USA, 61 pp.
  849. Thornton, P.K., J. van de Steeg, A. Notenbaert, and M. Herrero, 2009: The impacts of climate change on livestock and livestock systems in developing countries: A review of what we know and what we need to know. Agric. Syst., 101, 113–127, doi:10.1016/J.AGSY.2009.05.002.
  850. Sherwood, S., and Q. Fu, 2014: A drier future? Science, 343, 737–739, doi:10.1126/science.1247620.
  851. Cook, B.I., J.E. Smerdon, R. Seager, and S. Coats, 2014a: Global warming and 21st century drying. Clim. Dyn., 43, 2607–2627, doi:10.1007/s00382-014-2075-y.
  852. Dai, A., 2011: Drought under global warming: A review. Wiley Interdiscip. Rev. Clim. Chang., 2, 45–65, doi:10.1002/wcc.81.
  853. Dominguez, F., J. Cañon, and J. Valdes, 2010: IPCC-AR4 climate simulations for the south-western US: The importance of future ENSO projections. Clim. Change, 99, 499–514, doi:10.1007/s10584-009-9672-5.
  854. Feng, S., and Q. Fu, 2013: Expansion of global drylands under a warming climate. Atmos. Chem. Phys., 13, 10081–10094, doi:10.5194/acp-13-10081-2013.
  855. Ficklin, D.L., J.T. Abatzoglou, S.M. Robeson, and A. Dufficy, 2016: The influence of climate model biases on projections of aridity and drought. J. Clim., 29, 1369–1389, doi:10.1175/JCLI-D-15-0439.1.
  856. Fu, Q., L. Lin, J. Huang, S. Feng, and A. Gettelman, 2016: Changes in terrestrial aridity for the period 850–2080 from the community earth system model. J. Geophys. Res., 121, 2857–2873, doi:10.1002/2015JD024075.
  857. Greve, P., and S.I. Seneviratne, 1999: Assessment of future changes in water availability and aridity. Geophys. Res. Lett., 42, 5493–5499, doi:10.1002/2015GL064127.
  858. Koutroulis, A.G., 2019: Dryland changes under different levels of global warming. Sci. Total Environ., 655, 482–511, doi:10.1016/J.SCITOTENV.2018.11.215.
  859. Scheff, J., and D.M. W Frierson, 2015: Terrestrial aridity and its response to greenhouse warming across CMIP5 climate models. J. Clim., 28, 5583–5600, doi:10.1175/JCLI-D-14-00480.1.
  860. Zhao, S., H. Zhang, S. Feng, and Q. Fu, 2015: Simulating direct effects of dust aerosol on arid and semi-arid regions using an aerosol-climate coupled system. Int. J. Climatol., 35, 1858–1866, doi:10.1002/joc.4093.
  861. Zeng, N., and J. Yoon, 2009: Expansion of the world’s deserts due to vegetation-albedo feedback under global warming. Geophys. Res. Lett., 36, L17401, doi:10.1029/2009GL039699.
  862. Cook, B.I., J.E. Smerdon, R. Seager, and S. Coats, 2014a: Global warming and 21st century drying. Clim. Dyn., 43, 2607–2627, doi:10.1007/s00382-014-2075-y.
  863. Fu, Q., L. Lin, J. Huang, S. Feng, and A. Gettelman, 2016: Changes in terrestrial aridity for the period 850–2080 from the community earth system model. J. Geophys. Res., 121, 2857–2873, doi:10.1002/2015JD024075.
  864. Lin, L., A. Gettelman, S. Feng, and Q. Fu, 2015: Simulated climatology and evolution of aridity in the 21st century. J. Geophys. Res. Atmos., 120, 5795–5815, doi:10.1002/2014JD022912.
  865. Scheff, J., and D.M. W Frierson, 2015: Terrestrial aridity and its response to greenhouse warming across CMIP5 climate models. J. Clim., 28, 5583–5600, doi:10.1175/JCLI-D-14-00480.1.
  866. Roderick, M.L., P. Greve, and G.D. Farquhar, 2015: On the assessment of aridity with changes in atmospheric CO2. Water Resour. Res., 51, 5450–5463, doi:10.1002/2015WR017031.
  867. Fu, Q., C.M. Johanson, J.M. Wallace, and T. Reichler, 2006: Enhanced mid-latitude tropospheric warming in satellite measurements. Science, 312, 1179, doi:10.1126/science.1125566.
  868. Hu, Y., and Q. Fu, 2007: Observed poleward expansion of the Hadley circulation since 1979. Atmos. Chem. Phys., 7, 5229–5236, doi:10.5194/acp-7-5229-2007.
  869. Johanson, C.M., Q. Fu, C.M. Johanson, and Q. Fu, 2009: Hadley cell widening: Model simulations versus observations. J. Clim., 22, 2713–2725, doi:10.1175/2008JCLI2620.1.
  870. Seidel, D.J., and W.J. Randel, 2007: Recent widening of the tropical belt: Evidence from tropopause observations. J. Geophys. Res., 112, D20113, doi:10.1029/2007JD008861.
  871. Johanson, C.M., Q. Fu, C.M. Johanson, and Q. Fu, 2009: Hadley cell widening: Model simulations versus observations. J. Clim., 22, 2713–2725, doi:10.1175/2008JCLI2620.1.
  872. Lu, J., G.A. Vecchi, and T. Reichler, 2007: Expansion of the Hadley cell under global warming. Geophys. Res. Lett., 34, L06805, doi:10.1029/2006GL028443.
  873. Scheff, J., and D.M.W. Frierson, 2012: Robust future precipitation declines in CMIP5 largely reflect the poleward expansion of model subtropical dry zones. Geophys. Res. Lett., 39, doi:10.1029/2012GL052910.
  874. Terink, W., W.W. Immerzeel, and P. Droogers, 2013: Climate change projections of precipitation and reference evapotranspiration for the Middle East and northern Africa until 2050. Int. J. Climatol., 33, 3055–3072, doi:10.1002/joc.3650.
  875. Yin, Y., D. Ma, S. Wu, and T. Pan, 2015: Projections of aridity and its regional variability over China in the mid-21st century. Int. J. Climatol., 35, 4387–4398, doi:10.1002/joc.4295.
  876. Marengo, J.A., and M. Bernasconi, 2015: Regional differences in aridity/drought conditions over Northeast Brazil: Present state and future projections. Clim. Change, 129, 103–115, doi:10.1007/s10584-014-1310-1.
  877. Cook, B., N. Zeng, and J.H. Yoon, 2012: Will Amazonia dry out? Magnitude and causes of change from IPCC climate model projections. Earth Interact., 16, doi:10.1175/2011EI398.1.
  878. Nastos, P.T., N. Politi, and J. Kapsomenakis, 2013: Spatial and temporal variability of the Aridity Index in Greece. Atmos. Res., 119, 140–152, doi:10.1016/j.atmosres.2011.06.017.
  879. Coppola, E., and F. Giorgi, 2009: An assessment of temperature and precipitation change projections over Italy from recent global and regional climate model simulations. Int. J. Climatol., 30: 11-32, doi:10.1002/joc.1867.
  880. IPCC, 2013: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, USA.
  881. Ramarao, M.V.S., R. Krishnan, J. Sanjay, and T.P. Sabin, 2015: Understanding land surface response to changing South Asian monsoon in a warming climate. Earth Syst. Dyn., 6, 569–582, doi:10.5194/esd-6-569-2015.
  882. Hoegh-Guldberg, O., D. Jacob, M. Taylor, M. Bindi, S. Brown, I. Camilloni, A. Diedhiou, R. Djalante, K.L. Ebi, F. Engelbrecht, J.Guiot, Y. Hijioka, S. Mehrotra, A. Payne, S.I. Seneviratne, A. Thomas, R. Warren, and G. Zhou, 2018: Impacts of 1.5°C of Global Warming on Natural and Human Systems. In: Global Warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty [Masson-Delmotte, V., P. Zhai, 
H.-O. Pörtner, D. Roberts, J. Skea, P.R. Shukla, A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, J.B.R. Matthews, Y. Chen, X. Zhou, M.I.Gomis, E. Lonnoy, T.Maycock, M.Tignor, and T. Waterfield (eds.)]. In press. pp. 175–311.
  883. Miao, L., P. Ye, B. He, L. Chen, and X. Cui, 2015b: Future climate impact on the desertification in the Dryland Asia using AVHRR GIMMS NDVI3g data. Remote Sens., 7, 3863–3877, doi:10.3390/rs70403863.
  884. Schewe, J., and A. Levermann, 2017: Non-linear intensification of Sahel rainfall as a possible dynamic response to future warming. Earth Syst. Dynam, 8, 495–505, doi:10.5194/esd-8-495-2017.
  885. Evan, A.T., 2018: Surface winds and dust biases in climate models. Geophys. Res. Lett., 45, 1079–1085, doi:10.1002/2017GL076353.
  886. Evan, A.T., C. Flamant, S. Fiedler, and O. Doherty, 2014: An analysis of aeolian dust in climate models. Geophys. Res. Lett., 41, 5996–6001, doi:10.1002/2014GL060545.
  887. Mahowald, N.M. et al., 2006: Change in atmospheric mineral aerosols in response to climate: Last glacial period, preindustrial, modern, and doubled carbon dioxide climates. J. Geophys. Res. Atmos., 111:1-22, D10202, n/a–n/a, doi:10.1029/2005JD006653.
  888. Mahowald, N.M., 2007: Anthropocene changes in desert area: Sensitivity to climate model predictions. Geophys. Res. Lett., 34, L18817, doi:10.1029/2007GL030472.
  889. Wang, X., Y. Yang, Z. Dong, and C. Zhang, 2009: Responses of dune activity and desertification in China to global warming in the twenty-first century. Glob. Planet. Change, 67, 167–185, doi:10.1016/j.gloplacha.2009.02.004.
  890. Pu, B., and P. Ginoux, 2017: Projection of American dustiness in the late 21st century due to climate change. Sci. Rep., 7, 1–10, doi:10.1038/s41598-017-05431-9.
  891. Evan, A.T., C. Flamant, M. Gaetani, and F. Guichard, 2016: The past, present and future of African dust. Nature, 531, 493–495, doi:10.1038/nature17149.
  892. Leggett, J., W.J. Pepper, R.J. Swart, J. Edmonds, L.G. Meira Filho, I. Mintzer, M.X. Wang, and J. Wasson, 1992: Emissions scenarios for the IPCC: An update. In: Climate Change 1992: The Supplementary Report to IPCC Scientific Assessment [Houghton, J.T., B.A. Callander and S.K. Varney (eds.)]. Cambridge University Press, Cambridge, United Kingdom, pp. 69–95, doi:10.1007/s00213-003-1546-3.
  893. Schofield, R. V, and M.J. Kirkby, 2003: Application of salinization indicators and initial development of potential global soil salinization scenario under climatic change. Global Biogeochem. Cycles, 17, doi:10.1029/2002GB001935.
  894. Rengasamy, P., 2006: World salinization with emphasis on Australia. J. Exp. Bot., 57, 1017–1023, doi:10.1093/jxb/erj108.
  895. Sivakumar, M.V.K., Ndiang’ui, N. (ed.), 2007: Climate and Land Degradation (Environmental Science and Engineering). Springer-Verlag, Berlin, Germany, 623 pp.
  896. Sherif, M.M., and V.P. Singh, 1999: Effect of climate change on seawater intrusion in coastal aquifers. Hydrol. Process., 13, 1277–1287, doi:10.1002/(SICI)1099-1085(19990615)13:8<1277::AID-HYP765>3.0.CO;2-W.
  897. Nielsen, D.L., and M.A. Brock, 2009: Modified water regime and salinity as a consequence of climate change: Prospects for wetlands of southern Australia. Clim. Change, 95, 523–533, doi:10.1007/s10584-009-9564-8.
  898. IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA. ISBN 978-1-107-02506-6, 582 pp.
  899. Sherwood, S., and Q. Fu, 2014: A drier future? Science, 343, 737–739, doi:10.1126/science.1247620.
  900. IPCC, 2013: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, USA.
  901. Carrão, H., G. Naumann, and P. Barbosa, 2017: Global projections of drought hazard in a warming climate: A prime for disaster risk management. Clim. Dyn., 50, 2137–2155, doi:10.1007/s00382-017-3740-8.
  902. Morales, C. et al., 2011: Measuring the Economic Value of Land Degradation/Desertification Considering the Effects of Climate Change. A Study for Latin America and the Caribbean. Communication au Séminaire, Politiques, programmes et projets de lutte contre la désertification, quelles évaluations, Montpellier, France. 20 pages.
  903. Alessandri, A. et al., 2014: Robust assessment of the expansion and retreat of Mediterranean climate in the 21st century. Sci. Rep., 4, 4–11, doi:10.1038/srep07211.
  904. Polade, S.D., A. Gershunov, D.R. Cayan, M.D. Dettinger, and D.W. Pierce, 2017: Precipitation in a warming world: Assessing projected hydro-climate changes in California and other Mediterranean climate regions. Sci. Rep., 7, 1–10, doi:10.1038/s41598-017-11285-y.
  905. Alessandri, A. et al., 2014: Robust assessment of the expansion and retreat of Mediterranean climate in the 21st century. Sci. Rep., 4, 4–11, doi:10.1038/srep07211.
  906. Dai, A., 2013: Increasing drought under global warming in observations and models. Nat. Clim. Chang., 3, 52–58, doi:10.1038/nclimate1633.
  907. Sheffield, J., and E.F. Wood, 2008: Projected changes in drought occurrence under future global warming from multi-model, multi-scenario, IPCC AR4 simulations. Clim. Dyn., 31, 79–105, doi:10.1007/s00382-007-0340-z.
  908. Swann, A.L.S., F.M. Hoffman, C.D. Koven, and J.T. Randerson, 2016: Plant responses to increasing CO2 reduce estimates of climate impacts on drought severity. Proc. Natl. Acad. Sci. U.S.A., 113, 10019–10024, doi:10.1073/pnas.1604581113.
  909. Wang, G., 2005: Agricultural drought in a future climate: Results from 15 global climate models participating in the IPCC 4th assessment. Clim. Dyn., 25, 739–753, doi:10.1007/s00382-005-0057-9.
  910. Zhao, S., H. Zhang, S. Feng, and Q. Fu, 2015: Simulating direct effects of dust aerosol on arid and semi-arid regions using an aerosol-climate coupled system. Int. J. Climatol., 35, 1858–1866, doi:10.1002/joc.4093.
  911. Carrão, H., G. Naumann, and P. Barbosa, 2017: Global projections of drought hazard in a warming climate: A prime for disaster risk management. Clim. Dyn., 50, 2137–2155, doi:10.1007/s00382-017-3740-8.
  912. Naumann, G. et al., 2018: Global changes in drought conditions under different levels of warming. Geophys. Res. Lett., 45, 3285–3296, doi:10.1002/2017GL076521.
  913. Ukkola, A.M. et al., 2018: Evaluating CMIP5 model agreement for multiple drought metrics. J. Hydrometeorol., 19, 969–988, doi:10.1175/JHM-D-17-0099.1.
  914. Reyer, C.P.O. et al., 2013: A plant’s perspective of extremes: Terrestrial plant responses to changing climatic variability. Glob. Chang. Biol., 19, 75–89, doi:10.1111/gcb.12023.
  915. Li, Y. et al., 2018b: Climate model shows large-scale wind and solar farms in the Sahara increase rain and vegetation. Science, 361, 1019–1022, doi:10.1126/science.aar5629.
  916. Seddon, A.W.R., M. Macias-Fauria, P.R. Long, D. Benz, and K.J. Willis, 2016: Sensitivity of global terrestrial ecosystems to climate variability. Nature, 531, 229–232, doi:10.1038/nature16986.
  917. You, N., J. Meng, and L. Zhu, 2018: Sensitivity and resilience of ecosystems to climate variability in the semi-arid to hyper-arid areas of Northern China: A case study in the Heihe River Basin. Ecol. Res., 33, 161–174, doi:10.1007/s11284-017-1543-3.
  918. Sloat, L.L. et al., 2018: Increasing importance of precipitation variability on global livestock grazing lands. Nat. Clim. Chang., 8, 214–218, doi:10.1038/s41558-018-0081-5.
  919. Dougill, A.J., E.D.G. Fraser, and M.S. Reed, 2010: Anticipating vulnerability to climate change in Dryland pastoral systems: Using dynamic systems models for the Kalahari. Ecol. Soc., 15 (2): 17. http://www.ecologyandsociety.org/vol15/iss2/art17/.
  920. López-i-Gelats, F., E.D.G. Fraser, J.F. Morton, and M.G. Rivera-Ferre, 2016: What drives the vulnerability of pastoralists to global environmental change? A qualitative meta-analysis. Glob. Environ. Chang., 39, 258–274, doi:10.1016/J.GLOENVCHA.2016.05.011.
  921. Máñez Costa, M.A., E.J. Moors, and E.D.G. Fraser, 2011: Socio-economics, policy, or climate change: What is driving vulnerability in southern Portugal? Ecol. Soc., 16 (1): 28, doi:10.5751/ES-03703-160128.
  922. Liu, H.-L., P. Willems, A.-M. Bao, L. Wang, and X. Chen, 2016b: Effect of climate change on the vulnerability of a socio-ecological system in an arid area. Glob. Planet. Change, 137, 1–9, doi:10.1016/J.GLOPLACHA.2015.12.014.
  923. Byers, E. et al., 2018: Global exposure and vulnerability to multi-sector development and climate change hotspots. Environ. Res. Lett., 13, 55012, doi:10.1088/1748-9326/aabf45.
  924. Yang, D., S. Kanae, T. Oki, T. Koike, and K. Musiake, 2003: Global potential soil erosion with reference to land use and climate changes. Hydrol. Process., 17, 2913–2928, doi:10.1002/hyp.1441.
  925. Zare, M., A.A. Nazari Samani, M. Mohammady, T. Teimurian, and J. Bazrafshan, 2016: Simulation of soil erosion under the influence of climate change scenarios. Environ. Earth Sci., 75, 1–15, doi:10.1007/s12665-016-6180-6.
  926. Türkeş, M., M. Turp, T. An, N. Ozturk, and M.L. Kurnaz, 2019: Impacts of climate change on precipitation climatology and variability in Turkey. In: Water Resources of Turkey [Harmancioglu, N.B. and D. Altinbilek, (eds.)]. Springer International Publishing, New York, USA. Pages 467–491.
  927. Serpa, D. et al., 2015: Impacts of climate and land use changes on the hydrological and erosion processes of two contrasting Mediterranean catchments. Sci. Total Environ., 538, 64–77, doi:10.1016/j.scitotenv.
2015.08.033.
  928. Simonneaux, V., A. Cheggour, C. Deschamps, F. Mouillot, O. Cerdan, and 
Y. Le Bissonnais, 2015: Land use and climate change effects on soil erosion in a semi-arid mountainous watershed (High Atlas, Morocco). J. Arid Environ., 122, 64–75, doi:10.1016/j.jaridenv.2015.06.002.
  929. Jiménez Cisneros, B.E., T. Oki, N.W. Arnell, G. Benito, J.G. Cogley, P. Döll, 
T. Jiang, and S.S. Mwakalila, 2014: Freshwater resources. In: Climate Change 2014: Impacts,Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White (eds.)]. Cambridge University Press, New York, USA, pp. 229–269.
  930. Ravi, S., D.D. D Breshears, T.E. Huxman, P. D’Odorico,2010: Land degradation in drylands: Interactions among hydrologic-aeolian erosion and vegetation dynamics. Geomorphology, 116, 236–245, doi:10.1016/J.GEOMORPH.2009.11.023.
  931. Zabaleta, A., M. Meaurio, E. Ruiz, and I. Antigüedad, 2013: Simulation climate change impact on runoff and sediment yield in a small watershed in the Basque country, northern Spain. J. Environ. Qual., 43, 235, doi:10.2134/jeq2012.0209.
  932. Nerantzaki, S.D. et al., 2015: Modeling suspended sediment transport and assessing the impacts of climate change in a karstic Mediterranean watershed. Sci. Total Environ., 538, 288–297, doi:10.1016/j.scitotenv.2015.07.092.
  933. Klik, A., and J. Eitzinger, 2010: Impact of climate change on soil erosion and the efficiency of soil conservation practices in Austria. J. Agric. Sci., 148, 529–541, doi:10.1017/S0021859610000158.
  934. Borrelli, P. et al., 2017: An assessment of the global impact of 21st century land use change on soil erosion. Nat. Commun., 8, 2013, doi:10.1038/s41467-017-02142-7.
  935. Huang, J., H. Yu, A. Dai, Y. Wei, and L. Kang, 2017: Drylands face potential threat under 2°C global warming target. Nat. Clim. Chang., 7, 417–422, doi:10.1038/nclimate3275.
  936. Yi, C., S. Wei, and G. Hendrey, 2014: Warming climate extends dryness-
controlled areas of terrestrial carbon sequestration. Sci. Rep., 4, 1–6, doi:10.1038/srep05472.
  937. Ivanov, A.L. et al., 2018: National report global climate and soil cover of Russia: Assessment of risks and environmental and economic consequences of land degradation. Adaptive systems and technologies of environmental management (agriculture and forestry). In: Russian: NATSIONAL’N. Moscow, Russia, 357 pp.
  938. Rodriguez-Caballero, E., J. Belnap, B. Büdel, P.J. Crutzen, M.O. Andreae, U. Pöschl, and B. Weber, 2018: Dryland photoautotrophic soil surface communities endangered by global change. Nat. Geosci., 11, 185–189, doi:10.1038/s41561-018-0072-1.
  939. Rodriguez-Caballero, E., J. Belnap, B. Büdel, P.J. Crutzen, M.O. Andreae, U. Pöschl, and B. Weber, 2018: Dryland photoautotrophic soil surface communities endangered by global change. Nat. Geosci., 11, 185–189, doi:10.1038/s41561-018-0072-1.
  940. Rutherford, W.A., T.H. Painter, S. Ferrenberg, J. Belnap, G.S. Okin, C. Flagg, and S.C. Reed, 2017: Albedo feedbacks to future climate via climate change impacts on dryland biocrusts. Sci. Rep., 7, 44188, doi:10.1038/srep44188.
  941. Bachelet, D., K. Ferschweiler, T. Sheehan, and J. Strittholt, 2016: Climate change effects on southern California deserts. J. Arid Environ., 127, 17–29, doi:10.1016/J.JARIDENV.2015.10.003.
  942. García-Fayos, P., and E. Bochet, 2009: Indication of antagonistic interaction between climate change and erosion on plant species richness and soil properties in semi-arid Mediterranean ecosystems. Glob. Chang. Biol., 15, 306–318, doi:10.1111/j.1365-2486.2008.01738.x.
  943. Settele, J., R. Scholes, R. Betts, S. Bunn, P. Leadley, D. Nepstad, J.T. Overpeck, and M.A. Taboada, 2014: Terrestrial and inland water systems. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, 
M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, 
A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White, (eds.)]. Cambridge University Press, New York, USA, pp. 271–359.
  944. Malcolm, J.R., C. Liu, R.P. Neilson, L. Hansen, and L. Hannah, 2006: Global warming and extinctions of endemic species from biodiversity hotspots. Conserv. Biol., 20, 538–548, doi:10.1111/j.1523-1739.2006.00364.x.
  945. van Vuuren, D.P., O.E. Sala, and H.M. Pereira, 2006: The future of vascular plant diversity under four global scenarios. Ecol. Soc., 11 (2): 5, 25. Van Vuuren, D. P., O. E. Sala, and H. M. Pereira. 2006. The future of vascular plant diversity under four global scenarios. Ecology and Society 11(2): 25. [online] URL: http://www.ecologyandsociety.org/vol11/iss2/art25/.
  946. Ye, X., X. Yu, C. Yu, A. Tayibazhaer, F. Xu, A.K. Skidmore, and T. Wang, 2018: Impacts of future climate and land cover changes on threatened mammals in the semi-arid Chinese Altai Mountains. Sci. Total Environ., 612, 775–787, doi:10.1016/J.SCITOTENV.2017.08.191.
  947. Fischlin, A., G.F. Midgley, J. Price, R. Leemans, B. Gopal, C. Turley, M. Rounsevell, P. Dube, J. Tarazona, A. Velichko, 2007: Ecosystems, Their Properties, Goods, and Services. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. Van der Linden, and C.E. Hanson (eds.)]. 
Cambridge University Press, Cambridge, United Kingdom and New York, 
USA, 211–272 pp.
  948. Waha, K. et al., 2017: Climate change impacts in the Middle East and northern Africa (MENA) region and their implications for vulnerable population groups. Reg. Environ. Chang., 17, 1623–1638, doi:10.1007/s10113-017-1144-2.
  949. Betts, R.A. et al., 2018: Changes in climate extremes, fresh water availability and vulnerability to food insecurity projected at 1.5°C and 2°C global warming with a higher-resolution global climate model. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci., 376, 20160452, doi:10.1098/rsta.2016.0452.
  950. Serdeczny, O. et al., 2017: Climate change impacts in Sub-Saharan Africa: From physical changes to their social repercussions. Reg. Environ. Chang., 17, 1585–1600, doi:10.1007/s10113-015-0910-2.
  951. Wang, X., Y. Yang, Z. Dong, and C. Zhang, 2009: Responses of dune activity and desertification in China to global warming in the twenty-first century. Glob. Planet. Change, 67, 167–185, doi:10.1016/j.gloplacha.2009.02.004.
  952. Reyer, C.P.O., 2017: Climate change impacts in Central Asia and their implications for development. Reg. Environ. Chang., 17, 1639–1650, doi:10.1007/s10113-015-0893-z.
  953. World Bank, 2009: World Development Report 2010: Development and Climate Change. World Bank Group, Washington, DC, doi:10.1596/978-0-8213-7987-5, 417 pp.
  954. Knox, J., T. Hess, A. Daccache, and T. Wheeler, 2012: Climate change impacts on crop productivity in Africa and South Asia. Environ. Res. Lett., 7, 34032, doi:10.1088/1748-9326/7/3/034032.
  955. Schlenker, W., and D.B. Lobell, 2010: Robust negative impacts of climate change on African agriculture. Environ. Res. Lett., 5, 14010, doi:10.1088/1748-9326/5/1/014010.
  956. Nelson, G.C. et al., 2010: Food Security, Farming, and Climate Change to 2050: Scenarios, Results, Policy Options. International Food Policy Research Institute, Washington, DC, USA. DOI: 10.2499/9780896291867, 131 pp.
  957. Hertel, T.W., M.B. Burke, and D.B. Lobell, 2010: The poverty implications of climate-induced crop yield changes by 2030. Glob. Environ. Chang., 20, 577–585, doi:10.1016/J.GLOENVCHA.2010.07.001.
  958. Ivanov, A.L. et al., 2018: National report global climate and soil cover of Russia: Assessment of risks and environmental and economic consequences of land degradation. Adaptive systems and technologies of environmental management (agriculture and forestry). In: Russian: NATSIONAL’N. Moscow, Russia, 357 pp.
  959. Hallegatte, S., and J. Rozenberg, 2017: Climate change through a poverty lens. Nat. Clim. Chang., 7, 250–256, doi:10.1038/nclimate3253.
  960. Hertel, T.W., and D.B. Lobell, 2014: Agricultural adaptation to climate change in rich and poor countries: Current modeling practice and potential for empirical contributions. Energy Econ., 46, 562–575, doi:10.1016/J.ENECO.2014.04.014.
  961. Hallegatte, S. et al., 2016: Shock Waves: Managing the Impacts of Climate Change on Poverty. The World Bank, Washington DC, USA, 207 pp.
  962. Hallegatte, S. et al., 2016: Shock Waves: Managing the Impacts of Climate Change on Poverty. The World Bank, Washington DC, USA, 207 pp.
  963. Hertel, T.W., M.B. Burke, and D.B. Lobell, 2010: The poverty implications of climate-induced crop yield changes by 2030. Glob. Environ. Chang., 20, 577–585, doi:10.1016/J.GLOENVCHA.2010.07.001.
  964. Hertel, T.W., M.B. Burke, and D.B. Lobell, 2010: The poverty implications of climate-induced crop yield changes by 2030. Glob. Environ. Chang., 20, 577–585, doi:10.1016/J.GLOENVCHA.2010.07.001.
  965. Thomas, R., 2008: Opportunities to reduce the vulnerability of Dryland farmers in Central and West Asia and North Africa to climate change. Agric. Ecosyst. Environ., 126, 36–45, doi:10.1016/j.agee.2008.01.011.
  966. Scholes, R.J., 2009: Syndromes of dryland degradation in southern Africa. African J. Range Forage Sci., 26, 113–125, doi:10.2989/AJRF.2009.26.3.2.947.
  967. Liniger, H., and W. Critchley, 2007: Case Studies and Analysis of Soil and Water Conservation Initiatives Worldwide. University of Minnesota, Minnesota, USA, 364 pp.
  968. Scholes, R.J., 2009: Syndromes of dryland degradation in southern Africa. African J. Range Forage Sci., 26, 113–125, doi:10.2989/AJRF.2009.26.3.2.947.
  969. Reynolds, J.F. et al., 2007: Global desertification: Building a science for Dryland development. Science, 316, 847–851, doi:10.1126/science.1131634.
  970. Stringer, L.C. et al., 2017: A new Dryland development paradigm grounded in empirical analysis of Dryland systems science. L. Degrad. Dev., 28, 1952–1961, doi:10.1002/ldr.2716.
  971. Webb, N.P., N.A. Marshall, L.C. Stringer, M.S. Reed, A. Chappell, and J.E. Herrick, 2017: Land degradation and climate change: Building climate resilience in agriculture. Front. Ecol. Environ., 15, 450–459, doi:10.1002/fee.1530.
  972. de Coninck, H. et al., 2018: Chapter 4: Strengthening and Implementing the Global Response. In: Global Warming of 1.5°C. An IPCC Special Report on the Impacts of Global Warming of 1.5°C Above Pre-Industrial Levels and Related Global Greenhouse Gas Emission Pathways, in the Context of Strengthening the Global Response to the Threat of Climate Change, Sustainable Development, and Efforts to Eradicate Poverty [Masson-Delmotte, V., P. Zhai, H.O. Pörtner, D. Roberts, J. Skea, P.R. Shukla, 
A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, 
J.B.R. Matthews, Y. Chen, X. Zhou, M.I. Gomis, E. Lonnoy, T. Maycock, 
M. Tignor, T. Waterfield (eds.)]. In Press.
  973. IPBES, 2018: Assessment Report on Land Degradation and Restoration [Montanarella, L., R. Scholes., and A. Brainich. (eds.)]. Secretariate of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, Bonn, Germany, 686 pp.
  974. Webb, N.P., C.J. Stokes, and N.A. Marshall, 2013: Integrating biophysical and socio-economic evaluations to improve the efficacy of adaptation assessments for agriculture. Glob. Environ. Chang., 23, 1164–1177, doi:10.1016/j.gloenvcha.2013.04.007.
  975. Cowie, A.L. et al., 2018: Land in balance: The scientific conceptual framework for. Environ. Sci. Policy, 79, 25–35, doi:10.1016/J.ENVSCI.2017.10.011.
  976. Orr, B.J. et al., 2017: Scientific Conceptual Framework for Land Degradation Neutrality. A Report of the Science-Policy Interface. United Nations Convention to Combat Desertification, Bonn, Germany, 128 pp.
  977. Cherlet, M. et al. (eds.), 2018: World Atlas of Desertification. Publication Office of the European Union, Luxembourg, 248 pp.
  978. Knörzer, H., S. Graeff-Hönninger, B. Guo, P. Wang, and W. Claupein, 2009: The rediscovery of intercropping in China: A traditional cropping system for future Chinese agriculture – A review. In: Climate Change, Intercropping, Pest Control and Beneficial Microorganisms [Lichtfouse, E (ed.)]. Springer Netherlands, Dordrecht, Netherlands, pp. 13–44.
  979. Altieri, M.A., C.I. Nicholls, A. Henao, and M.A. Lana, 2015: Agroecology and the design of climate change-resilient farming systems. Agron. Sustain. Dev., 35, 869–890, doi:10.1007/s13593-015-0285-2.
  980. Amundson, R. et al., 2015: Soil science. Soil and human security in the 21st century. Science, 348, doi: 10.1126/science.1261071.
  981. Derpsch, R., T. Friedrich, A. Kassam, and L. Hongwen, 2010: Current status of adoption of no-till farming in the world and some of its main benefits. Int. J. Agric. Biol. Eng., 3, 1–25, doi:10.3965/j.issn.1934-6344.2010.01.0-0.
  982. Lal, R., 1997: Degradation and resilience of soils. Philos. Trans. R. Soc. London. Ser. B Biol. Sci., 352, 997–1010, doi:10.1098/rstb.1997.0078.
  983. de Vries, F.T., M.E. Liiri, L. Bjørnlund, M. Bowker, S. Christensen, H. Setälä, and R.D. Bardgett, 2012: Land use alters the resistance and resilience of soil food webs to drought. Nat. Clim. Chang., 2, 276–280, doi:10.1038/nclimate1368.
  984. Touré, A., et al., 2019: Dynamics of wind erosion and impact of vegetation cover and land use in the Sahel: A case study on sandy dunes in south-eastern Niger. Catene, 177, 272–285, doi: 10.1016/j.catena.2019.02.011.
  985. Amundson, R. et al., 2015: Soil science. Soil and human security in the 21st century. Science, 348, doi: 10.1126/science.1261071.
  986. Borrelli, P. et al., 2017: An assessment of the global impact of 21st century land use change on soil erosion. Nat. Commun., 8, 2013, doi:10.1038/s41467-017-02142-7.
  987. Pierre, C. et al., 2017: Impact of agropastoral management on wind erosion in Sahelian croplands. L. Degrad. Dev., 29, 800–811, doi:10.1002/ldr.2783.
  988. Amadou, A. et al., 2011: Impact of very low crop residues cover on wind erosion in the Sahel. Catena, 85, 205–214, doi:10.1016/j.catena.2011.01.002.
  989. Ibragimov, N. et al., 2007: Water use efficiency of irrigated cotton in Uzbekistan under drip and furrow irrigation. Agric. Water Manag., 90, 112–120, doi:10.1016/j.agwat.2007.01.016.
  990. Narayanamoorthy, A., 2010: Can drip method of irrigation be used to achieve the macro objectives of conservation agriculture? Indian J. Agric. Econ., 65, 428–438.
  991. Niaz, S., M.A. Ali, S. Ali, and S. Awan, 2009: Comparative water use efficiency of drip and furrow irrigation systems for off-season vegetables under plastic tunnel in rainfed areas. Life Sci. Int. J., 2, 952–955.
  992. Niaz, S., M.A. Ali, S. Ali, and S. Awan, 2009: Comparative water use efficiency of drip and furrow irrigation systems for off-season vegetables under plastic tunnel in rainfed areas. Life Sci. Int. J., 2, 952–955.
  993. Narayanamoorthy, A., 2010: Can drip method of irrigation be used to achieve the macro objectives of conservation agriculture? Indian J. Agric. Econ., 65, 428–438.
  994. Ibragimov, N. et al., 2007: Water use efficiency of irrigated cotton in Uzbekistan under drip and furrow irrigation. Agric. Water Manag., 90, 112–120, doi:10.1016/j.agwat.2007.01.016.
  995. Altieri, M., and P. Koohafkan, 2008: Enduring Farms: Climate Change, Smallholders and Traditional Farming Communities. Third World Network, TWN Environment & Development Series no. 6, Penang, Malaysia. ISBN: 978-983-2729-55-6, 63 pp.
  996. Tanveer, M., S.A. Anjum, S. Hussain, A. Cerdà, and U. Ashraf, 2017: Relay cropping as a sustainable approach: Problems and opportunities for sustainable crop production. Environ. Sci. Pollut. Res., 24, 6973–6988, doi:10.1007/s11356-017-8371-4.
  997. Wilhelm, W.W., and C.S. Wortmann, 2004: Tillage and rotation interactions for corn and soybean grain yield as affected by precipitation and air temperature. Agron. J., 96, 425–432, doi:10.2134/agronj2004.4250.
  998. Khan, Z.R. et al., 2014: Achieving food security for one million Sub-Saharan African poor through push-pull innovation by 2020. Philos. Trans. R. Soc. B Biol. Sci., 369, 20120284, doi:10.1098/rstb.2012.0284.
  999. Wang, Y. et al., 2008: The dynamics variation of soil moisture of shelterbelts along the Tarim Desert Highway. Sci. Bull., 53, 102–108, doi:10.1007/s11434-008-6011-6.
  1000. Wang, Y. et al., 2008: The dynamics variation of soil moisture of shelterbelts along the Tarim Desert Highway. Sci. Bull., 53, 102–108, doi:10.1007/s11434-008-6011-6.
  1001. Nyamadzawo, G., M. Wuta, J. Nyamangara, and D. Gumbo, 2013: Opportunities for optimization of in-field water harvesting to cope with changing climate in semi-arid smallholder farming areas of Zimbabwe. Springerplus, 2, 1–9, doi:10.1186/2193-1801-2-100.
  1002. Nyagumbo, I., G. Nyamadzawo, and C. Madembo, 2019: Effects of three in-field water harvesting technologies on soil water content and maize yields in a semi-arid region of Zimbabwe. Agric. Water Manag., 216, 206–213, doi:10.1016/J.AGWAT.2019.02.023.
  1003. Garrity, D.P. et al., 2010: Evergreen agriculture: A robust approach to sustainable food security in Africa. Food Secur., 2, 197–214, doi:10.1007/s12571-010-0070-7.
  1004. Taye, G. et al., 2015: Evolution of the effectiveness of stone bunds and trenches in reducing runoff and soil loss in the semi-arid Ethiopian highlands. Zeitschrift für Geomorphol., 59, 477–493, doi:10.1127/zfg/2015/0166.
  1005. Dile, Y.T., L. Karlberg, M. Temesgen, and J. Rockström, 2013: The role of water harvesting to achieve sustainable agricultural intensification and resilience against water-related shocks in Sub-Saharan Africa. Agric. Ecosyst. Environ., 181, 69–79, doi:10.1016/j.agee.2013.09.014.
  1006. Vohland, K., and B. Barry, 2009: A review of in situ rainwater harvesting (RWH) practices modifying landscape functions in African drylands. Agric. Ecosyst. Environ., 131, 119–127, doi:10.1016/j.agee.2009.01.010.
  1007. Almazroui, M., M.N. Islam, K.S. Balkhair, Z. Şen, and A. Masood, 2017: Rainwater harvesting possibility under climate change: A basin-scale case study over western province of Saudi Arabia. Atmos. Res., 189, 11–23, doi:10.1016/J.ATMOSRES.2017.01.004.
  1008. Biazin, B., G. Sterk, M. Temesgen, A. Abdulkedir, and L. Stroosnijder, 2012: Rainwater harvesting and management in rainfed agricultural systems in Sub-Saharan Africa – A review. Phys. Chem. Earth, Parts A/B/C, 47–48, 139–151, doi:10.1016/J.PCE.2011.08.015.
  1009. Bouma, J., and J.H.M. Wösten, 2016: How to characterize ‘good’ and ‘greening’ in the EU Common Agricultural Policy (CAP): The case of clay soils in the Netherlands. Soil Use Manag., 32, 546–552, doi:10.1111/sum.12289.
  1010. Dile, Y.T., L. Karlberg, M. Temesgen, and J. Rockström, 2013: The role of water harvesting to achieve sustainable agricultural intensification and resilience against water-related shocks in Sub-Saharan Africa. Agric. Ecosyst. Environ., 181, 69–79, doi:10.1016/j.agee.2013.09.014.
  1011. Bouma, J., and J.H.M. Wösten, 2016: How to characterize ‘good’ and ‘greening’ in the EU Common Agricultural Policy (CAP): The case of clay soils in the Netherlands. Soil Use Manag., 32, 546–552, doi:10.1111/sum.12289.
  1012. Bouma, J., and J.H.M. Wösten, 2016: How to characterize ‘good’ and ‘greening’ in the EU Common Agricultural Policy (CAP): The case of clay soils in the Netherlands. Soil Use Manag., 32, 546–552, doi:10.1111/sum.12289.
  1013. Vohland, K., and B. Barry, 2009: A review of in situ rainwater harvesting (RWH) practices modifying landscape functions in African drylands. Agric. Ecosyst. Environ., 131, 119–127, doi:10.1016/j.agee.2009.01.010.
  1014. Meijer, E., E. Querner, and H. Boesveld, 2013: Impact of Farm Dams on River Flows; A Case Study in the Limpopo River Basin, Southern Africa. Alterra report 2394, Alterra Wageningen UR, Wageningen, Netherlands. 61 pages.
  1015. Singh, G. et al., 2012: Effects of rainwater harvesting and afforestation on soil properties and growth of Emblica officinalis while restoring degraded hills in western India. African J. Environ. Sci. Technol., 6, 300–311, doi:10.5897/AJEST11.040.
  1016. Vohland, K., and B. Barry, 2009: A review of in situ rainwater harvesting (RWH) practices modifying landscape functions in African drylands. Agric. Ecosyst. Environ., 131, 119–127, doi:10.1016/j.agee.2009.01.010.
  1017. Yosef, B.A.A., and D.K.K. Asmamaw, 2015: Rainwater harvesting: An option for dry land agriculture in arid and semi-arid Ethiopia. Int. J. Water Resour. Environ. Eng., 7, 17–28, doi:10.5897/IJWREE2014.0539.
  1018. Boelee, E. et al., 2013: Options for water storage and rainwater harvesting to improve health and resilience against climate change in Africa. Reg. Environ. Chang., 13, 509–519, doi:10.1007/s10113-012-0287-4.
  1019. Lal, R., 2004: Carbon Sequestration in Dryland Ecosystems. Environ. Manage., 33 (4), 528–544, doi:10.1007/s00267-003-9110-9.
  1020. Plaza-Bonilla, D. et al., 2015: Carbon management in dryland agricultural systems. A review. Agron. Sustain. Dev., 35, 1319–1334, doi:10.1007/s13593-015-0326-x.
  1021. Lal, R., 2015: The soil–peace nexus: Our common future. Soil Sci. Plant Nutr., 61, 566–578, doi:10.1080/00380768.2015.1065166.
  1022. Srinivasa Rao, C.H., K. Sumantra, and P. Thakur, 2015: Carbon sequestration through conservation agriculture in rainfed systems. In: Integrated Soil and Water Resource Management for Livelihood and Environmental Security [Rajkhowa, D.J., A. Das, S.V. Ngachan, A. Sikka, and M. Lyngdoh (eds.)]. ICAR Research Complexfor NEH Region, Umiam, Meghalaya, India, pp. 56–67.
  1023. Sombrero, A., and A. De Benito, 2010: Carbon accumulation in soil. Ten-year study of conservation tillage and crop rotation in a semi-arid area of Castile-Leon, Spain. Soil Tillage Res., 107, 64–70, doi:10.1016/j.still.2010.02.009.
  1024. Cheesman, S., C. Thierfelder, N.S. Eash, G.T. Kassie, and E. Frossard, 2016: Soil carbon stocks in conservation agriculture systems of southern Africa. Soil Tillage Res., 156, 99–109, doi:10.1016/j.still.2015.09.018.
  1025. Plaza-Bonilla, D. et al., 2015: Carbon management in dryland agricultural systems. A review. Agron. Sustain. Dev., 35, 1319–1334, doi:10.1007/s13593-015-0326-x.
  1026. Farage, P.K., J. Ardo, L. Olsson, E.A. Rienzi, A.S. Ball, and J.N. Pretty, 2007: The potential for soil carbon sequestration in three tropical Dryland farming systems of Africa and Latin America: A modelling approach. Soil Tillage Res., 94, 457–472, doi:10.1016/j.still.2006.09.006.
  1027. Hoyle, F.C., M.D. D’Antuono, T. Overheu, and D.V. Murphy, 2013: Capacity for increasing soil organic carbon stocks in Dryland agricultural systems. Soil Res., 51, 657–667.
  1028. Lal, R., 2004: Carbon Sequestration in Dryland Ecosystems. Environ. Manage., 33 (4), 528–544, doi:10.1007/s00267-003-9110-9.
  1029. Safriel, U., Adeel, Z. et al., 2005: Dryland systems. In: Ecosystems and Human Well-Being: Current State and Trends [Hassan, R., R.J. Scholes, and N. Ash, (eds.)]. Island Press, Washington, DC, USA, pp. 623–662.
  1030. Scholes, R.J., and S.R. Archer, 1997: Tree-grass interactions in savannas. Annu. Rev. Ecol. Syst., 28, 517–544, doi:10.1146/annurev.ecolsys.28.1.517.
  1031. Staver, A.C., S. Archibald, and S.A. Levin, 2011: The global extent and determinant of savanna and forest as alternative biome states. Science, 334, 230–233. DOI: 10.1126/science.1210465.
  1032. Stevens, N., C.E.R. Lehmann, B.P. Murphy, and G. Durigan, 2017: Savanna woody encroachment is widespread across three continents. Glob. Chang. Biol., 23, 235–244, doi:10.1111/gcb.13409.
  1033. Plaza-Bonilla, D. et al., 2015: Carbon management in dryland agricultural systems. A review. Agron. Sustain. Dev., 35, 1319–1334, doi:10.1007/s13593-015-0326-x.
  1034. Short, A.D., T.G. O’Connor, and C.R. Hurt, 2003: Medium-term changes in grass composition and diversity of Highland Sourveld grassland in the southern Drakensberg in response to fire and grazing management. African J. Range Forage Sci., 20, 1–10, doi:10.2989/10220110309485792.
  1035. Divinsky, I., N. Becker, and P. Bar (Kutiel), 2017: Ecosystem service tradeoff between grazing intensity and other services – A case study in Karei-Deshe experimental cattle range in northern Israel. Ecosyst. Serv., 24, 16–27, doi:10.1016/j.ecoser.2017.01.002.
  1036. Pellegrini, A.F.A. et al., 2017: Fire frequency drives decadal changes in soil carbon and nitrogen and ecosystem productivity. Nature, 553, 194–198, doi:10.1038/nature24668.
  1037. Devine, A.P., R.A. McDonald, T. Quaife, and I.M.D. Maclean, 2017: Determinants of woody encroachment and cover in African savannas. Oecologia, 183, 939–951, doi:10.1007/s00442-017-3807-6.
  1038. Donohue, R.J., M.L. Roderick, T.R. McVicar, and G.D. Farquhar, 2013: Impact of CO2 fertilization on maximum foliage cover across the globe’s warm, arid environments. Geophys. Res. Lett., 40, 3031–3035, doi:10.1002/grl.50563.
  1039. Divinsky, I., N. Becker, and P. Bar (Kutiel), 2017: Ecosystem service tradeoff between grazing intensity and other services – A case study in Karei-Deshe experimental cattle range in northern Israel. Ecosyst. Serv., 24, 16–27, doi:10.1016/j.ecoser.2017.01.002.
  1040. Fleischner, T.L., 1994: Costos ecológicos del pastoreo de ganado en el oeste de Estados Unidos. Conserv. Biol., 8, 629–644, doi:10.1046/j.1523-1739.1994.08030629.x.
  1041. van Oijen, M., G. Bellocchi, and M. Höglind, 2018: Effects of climate change on grassland biodiversity and productivity: The need for a diversity of models. Agronomy, 8, 14, doi:10.3390/agronomy8020014.
  1042. Plaza-Bonilla, D. et al., 2015: Carbon management in dryland agricultural systems. A review. Agron. Sustain. Dev., 35, 1319–1334, doi:10.1007/s13593-015-0326-x.
  1043. van Oijen, M., G. Bellocchi, and M. Höglind, 2018: Effects of climate change on grassland biodiversity and productivity: The need for a diversity of models. Agronomy, 8, 14, doi:10.3390/agronomy8020014.
  1044. O’Connor, T.G. et al., 2011: Influence of grazing management on plant diversity of Highland Sourveld influence of grazing management on plant diversity of Highland Sourveld grassland. Rangel. Ecol. Manag., 64, 196–207, doi:10.2111/REM-D-10-00062.1.
  1045. Scott-shaw, R., and C.D. Morris, 2015: Grazing depletes forb species diversity in the mesic grasslands of KwaZulu-Natal, South Africa. African J. Range Forage Sci., 32, 37–41, doi:10.2989/10220119.2014.901418.
  1046. Divinsky, I., N. Becker, and P. Bar (Kutiel), 2017: Ecosystem service tradeoff between grazing intensity and other services – A case study in Karei-Deshe experimental cattle range in northern Israel. Ecosyst. Serv., 24, 16–27, doi:10.1016/j.ecoser.2017.01.002.
  1047. Marty, J., 2005: Effects of cattle grazing on diversity in ephemeral wetlands. Conserv. Biol., 19, 1626–1632, doi:10.1111/j.1523-1739.2005.00198.x.
  1048. Plaza-Bonilla, D. et al., 2015: Carbon management in dryland agricultural systems. A review. Agron. Sustain. Dev., 35, 1319–1334, doi:10.1007/s13593-015-0326-x.
  1049. Han, G., X. Hao, M. Zhao, M. Wang, B.H. Ellert, W. Willms, and M. Wang, 2008: Effect of grazing intensity on carbon and nitrogen in soil and vegetation in a meadow steppe in Inner Mongolia. Agric. Ecosyst. Environ., 125, 21–32, doi:10.1016/j.agee.2007.11.009.
  1050. Bikila, N.G., Z.K. Tessema, and E.G. Abule, 2016: Carbon sequestration potentials of semi-arid rangelands under traditional management practices in Borana, southern Ethiopia. Agric. Ecosyst. Environ., 223, 108–114, doi:10.1016/J.AGEE.2016.02.028.
  1051. Manson, A.D., D. Jewitt, and A.D. Short, 2007: Effects of season and frequency of burning on soils and landscape functioning in a moist montane grassland. African J. Range Forage Sci., 24, 9–18.
  1052. Scholes, R.J., 2009: Syndromes of dryland degradation in southern Africa. African J. Range Forage Sci., 26, 113–125, doi:10.2989/AJRF.2009.26.3.2.947.
  1053. Papanastasis, V.P. et al., 2017: Comparative assessment of goods and services provided by grazing regulation and reforestation in degraded Mediterranean rangelands. L. Degrad. Dev., 28, 1178–1187, doi:10.1002/ldr.2368.
  1054. Scholes, R.J., 2009: Syndromes of dryland degradation in southern Africa. African J. Range Forage Sci., 26, 113–125, doi:10.2989/AJRF.2009.26.3.2.947.
  1055. Archer, S.R. et al., 2017: Woody plant encroachment: Causes and consequences. In: Rangeland Systems [Briske, D.D. (ed.)]. Springer Series on Environmental Management, Springer International Publishing, Cham, Switzerland, pp 25–84.
  1056. Conant, R.T., and K. Paustian, 2003: Potential soil carbon sequestration in overgrazed grassland ecosystems. Global Biogeochem. Cycles, 16, 90–99, doi:10.1029/2001gb001661.
  1057. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  1058. Geist, H.H.J., and E.F. Lambin, 2004: Dynamic Causal Patterns of Desertification. Bioscience, 54, 817–829, doi:10.1641/0006-3568(2004)054[0817:DCPOD]2.0.CO;2.
  1059. Havstad, K.M., L.F. Huenneke, and W.H. Schlesinger (eds.), 2006: Structure and Function of a Chihuahuan Desert Ecosystem: The Jornada Basin Long-Term Ecological Research Site. Oxford University Press, New York, USA, 247 pp.
  1060. Huang, D., K. Wang, and W.L. Wu, 2007: Dynamics of soil physical and chemical properties and vegetation succession characteristics during grassland desertification under sheep grazing in an agropastoral transition zone in northern China. J. Arid Environ., 70, 120–136, doi:10.1016/j.jaridenv.2006.12.009.
  1061. Manzano, M.G. and J. Návar, 2000: Processes of desertification by goats overgrazing in the Tamaulipan thornscrub (matorral) in north-eastern Mexico. J. Arid Environ., 44, 1–17, doi:10.1006/JARE.1999.0577.
  1062. Pointing, S.B., and J. Belnap, 2012: Microbial colonization and controls in Dryland systems. Nat. Rev. Microbiol., 10, 551–562, doi:10.1038/nrmicro2831.
  1063. Weber, B., M. Bowker, Y. Zhang, and J. Belnap, 2016: Natural recovery of biological soil crusts after disturbance. In: Biological Soil Crusts: An Organizing Principle in Drylands [Weber, B., B. Büdel, and J. Belnap (eds.)]. Springer International Publishing, New York, USA, pp. 479–498.
  1064. Smith, M.S., and B. Foran, 1992: An approach to assessing the economic risk of different drought management tactics on a South Australian pastoral sheep station. Agric. Syst., 39, 83–105.
  1065. Torell, L.A.A., S. Murugan, and O.A.A. Ramirez, 2010: Economics of flexible versus conservative stocking strategies to manage climate variability risk. Rangel. Ecol. Manag., 63, 415–425. doi.org/10.2111/REM-D-09-00131.1.
  1066. Khishigbayar, J. et al., 2015: Mongolian rangelands at a tipping point? Biomass and cover are stable but composition shifts and richness declines after 20 years of grazing and increasing temperatures. J. Arid Environ., 115, 100–112, doi:10.1016/j.jaridenv.2015.01.007.
  1067. Porensky, L.M., K.E. Mueller, D.J. Augustine, and J.D. Derner, 2016: Thresholds and gradients in a semi-arid grassland: Long-term grazing treatments induce slow, continuous and reversible vegetation change. J. Appl. Ecol., 53, 1013–1022, doi:10.1111/1365-2664.12630.
  1068. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  1069. Archer, S.R. et al., 2017: Woody plant encroachment: Causes and consequences. In: Rangeland Systems [Briske, D.D. (ed.)]. Springer Series on Environmental Management, Springer International Publishing, Cham, Switzerland, pp 25–84.
  1070. Abril, A., P. Barttfeld, and E.H. Bucher, 2005: The effect of fire and overgrazing disturbes on soil carbon balance in the Dry Chaco forest. For. Ecol. Manage., 206, 399–405, doi:10.1016/j.foreco.2004.11.014. doi.org/10.1016/j.foreco.2004.
11.014.
  1071. Abril, A., P. Barttfeld, and E.H. Bucher, 2005: The effect of fire and overgrazing disturbes on soil carbon balance in the Dry Chaco forest. For. Ecol. Manage., 206, 399–405, doi:10.1016/j.foreco.2004.11.014. doi.org/10.1016/j.foreco.2004.
11.014.
  1072. Bikila, N.G., Z.K. Tessema, and E.G. Abule, 2016: Carbon sequestration potentials of semi-arid rangelands under traditional management practices in Borana, southern Ethiopia. Agric. Ecosyst. Environ., 223, 108–114, doi:10.1016/J.AGEE.2016.02.028.
  1073. Bird, M.I., E.M. Veenendaal, C. Moyo, J. Lloyd, and P. Frost, 2000: Effect of fire and soil texture on soil carbon in a sub-humid savanna (Matopos, Zimbabwe). Geoderma, 94, 71–90, doi:10.1016/S0016-7061(99)00084-1.
  1074. Pellegrini, A.F.A. et al., 2017: Fire frequency drives decadal changes in soil carbon and nitrogen and ecosystem productivity. Nature, 553, 194–198, doi:10.1038/nature24668.
  1075. Abatzoglou, J.T., S.Z. Dobrowski, S.A. Parks, and K.C. Hegewisch, 2018: TerraClimate, a high-resolution global dataset of monthly climate and climatic water balance from 1958–2015. Sci. Data, 5, 170191, doi:10.1038/sdata.2017.191.
  1076. Littell, J., D. McKenzie, H.Y. Wan, and S. Cushman, 2018: Climate change and future wildfire in the western USA: An ecological approach to non-stationarity. Earth’s Futur., 6, 1097–1111, doi:10.1029/2018EF000878.
  1077. Moritz, M.A., M.-A. Parisien, E. Batllori, M.A. Krawchuk, J. Van Dorn, 
D.J. Ganz, and K. Hayhoe, 2012: Climate change and disruptions to global fire activity. Ecosphere, 3, art49, doi:10.1890/ES11-00345.1.
  1078. Fernandes, P.M. et al., 2013: Prescribed burning in southern Europe: Developing fire management in a dynamic landscape. Front. Ecol. Environ., 11, e4–e14, doi:10.1890/120298.
  1079. McCaw, W.L., 2013: Managing forest fuels using prescribed fire – A perspective from southern Australia. For. Ecol. Manage., 294, 217–224, doi:10.1016/j.foreco.2012.09.012.
  1080. van Wilgen, B.W. et al., 2010: Fire management in Mediterranean-climate shrublands: A case study from the Cape fynbos, South Africa. J. Appl. Ecol., 47, 631–638, doi:10.1111/j.1365-2664.2010.01800.x.
  1081. Webb, N.P., N.A. Marshall, L.C. Stringer, M.S. Reed, A. Chappell, and J.E. Herrick, 2017: Land degradation and climate change: Building climate resilience in agriculture. Front. Ecol. Environ., 15, 450–459, doi:10.1002/fee.1530.
  1082. Woollen, E. et al., 2016: Supplementary materials for charcoal production in the Mopane woodlands of Mozambique: What are the tradeoffs with other ecosystem services? Philos. Trans. R. Soc. B-Biological Sci., 371, 1–14, doi:10.1098/rstb.2015.0315.
  1083. Kloos, J., and F.G. Renaud, 2016: Overview of ecosystem-based approaches to drought risk reduction targeting small-scale farmers in Sub-Saharan Africa. In: Ecosystem Based Disaster Risk Reduction and Adaptation in Practice [Renaud, F.G., K. Sudmeier-Rieux, M. Estrella and U. Nehren (eds.)]. Springer, Cham, Switzerland, pp. 199–226.
  1084. Reid, H., A. Bourne, H. Muller, K. Podvin, S. Scorgie, and V. Orindi, 2018: Chapter 16 – A framework for assessing the effectiveness of ecosystem-based approaches to adaptation. In: Resilience – The Science of Adaptation to Climate Change [Zommers Z. and K. Alverson (eds.)]. Elsevier, Amsterdam, Netherlands, pp. 207–216.
  1085. Lal, R., 2004: Carbon Sequestration in Dryland Ecosystems. Environ. Manage., 33 (4), 528–544, doi:10.1007/s00267-003-9110-9.
  1086. Pelletier, J., A. Paquette, K. Mbindo, N. Zimba, A. Siampale, and B. Chendauka, 2018: Carbon sink despite large deforestation in African tropical dry forests (Miombo woodlands). Environ. Res. Lett., 13, 1–14, doi:10.1088/1748-9326/aadc9a.
  1087. Archer, S.R. et al., 2017: Woody plant encroachment: Causes and consequences. In: Rangeland Systems [Briske, D.D. (ed.)]. Springer Series on Environmental Management, Springer International Publishing, Cham, Switzerland, pp 25–84.
  1088. O’Connor, T. G, J.R. Puttick, and M.T. Hoffman, 2014: Bush encroachment in southern Africa: Changes and causes. African J. Range Forage Sci., 31, 67–88, doi:10.2989/10220119.2014.939996.
  1089. Schooley, R.L., B.T. Bestelmeyer, and A. Campanella, 2018: Shrub encroachment, productivity pulses, and core-transient dynamics of Chihuahuan Desert rodents. Ecosphere, 9, doi:10.1002/ecs2.2330.
  1090. Dougill, A.J. et al., 2016: Land use, rangeland degradation and ecological changes in the southern Kalahari, Botswana. Afr. J. Ecol., 54, 59–67, doi:10.1111/aje.12265.
  1091. O’Connor, T. G, J.R. Puttick, and M.T. Hoffman, 2014: Bush encroachment in southern Africa: Changes and causes. African J. Range Forage Sci., 31, 67–88, doi:10.2989/10220119.2014.939996.
  1092. Eldridge, D.J. et al., 2011: Impacts of shrub encroachment on ecosystem structure and functioning: Towards a global synthesis. Ecol. Lett., 14, 709–722, doi:10.1111/j.1461-0248.2011.01630.x.
  1093. Eldridge, D.J., and S. Soliveres, 2014: Are shrubs really a sign of declining ecosystem function? Disentangling the myths and truths of woody encroachment in Australia. Aust. J. Bot., 62, 594–608, doi:10.1071/BT14137.
  1094. Komala, P., and G. Prasad, 2016: Biomass: A key source of energy in rural households of Chamarajanagar district. Pelagia Res. Libr. Adv. Appl. Sci. Res., 7, 85–89.
  1095. Makonese, T., A. Ifegbesan, and I. Rampedi, 2017: Household cooking fuel use patterns and determinants across southern Africa: Evidence from the demographic and health survey data. Energy Environ., 29, 29–48, doi:10.1177/0958305X17739475.
  1096. Shackleton, C.M., and S.E. Shackleton, 2004: Use of woodland resources for direct household provisioning. In: Indigenous Forests and Woodlands in South Africa [Lawes, M.J., H.A.C. Eeley, C.M. Shackleton, and B.G.S. Geach (eds.)]. University of KwaZulu-Natal Press, Pietermaritzburg, South Africa.pp. 195–196.
  1097. Eldridge, D.J., and S. Soliveres, 2014: Are shrubs really a sign of declining ecosystem function? Disentangling the myths and truths of woody encroachment in Australia. Aust. J. Bot., 62, 594–608, doi:10.1071/BT14137.
  1098. Smit, G.N., J.N. de Klerk, M.B. Schneider, and J. van Eck, 2015: Detailed Assessment of the Biomass Resource and Potential Yield in a Selected Bush Encroached Area Of Namibia. Ministry of Agriculture, Water and Forestry, Windhoek, Namibia, 126 pp.
  1099. Stafford-Smith, M. et al., 2017: Integration: The Key to Implementing the Sustainable Development Goals. Sustain. Sci., 12, 911–919, doi:10.1007/s11625-016-0383-3.
  1100. Smit, G.N., J.N. de Klerk, M.B. Schneider, and J. van Eck, 2015: Detailed Assessment of the Biomass Resource and Potential Yield in a Selected Bush Encroached Area Of Namibia. Ministry of Agriculture, Water and Forestry, Windhoek, Namibia, 126 pp.
  1101. Smit, G.N., J.N. de Klerk, M.B. Schneider, and J. van Eck, 2015: Detailed Assessment of the Biomass Resource and Potential Yield in a Selected Bush Encroached Area Of Namibia. Ministry of Agriculture, Water and Forestry, Windhoek, Namibia, 126 pp.
  1102. Donohue, R.J., M.L. Roderick, T.R. McVicar, and G.D. Farquhar, 2013: Impact of CO2 fertilization on maximum foliage cover across the globe’s warm, arid environments. Geophys. Res. Lett., 40, 3031–3035, doi:10.1002/grl.50563.
  1103. Kgope, B.S., W.J. Bond, and G.F. Midgley, 2010: Growth responses of African savanna trees implicate atmospheric [CO2] as a driver of past and current changes in savanna tree cover. Austral Ecol., 35, 451–463, doi:10.1111/j.1442-9993.2009.02046.x.
  1104. Bond, W.J., and G.F. Midgley, 2000: A proposed CO2 controlled mechanism of woody plant invasion in grasslands and savannas. Glob. Chang. Biol., 6, 865–869.
  1105. Milton, S.J., 2003: ‘Emerging ecosystems’ – a washing-stone for ecologists, economists and sociologists? S. Afr. J. Sci., 99, 404–406.
  1106. Smit, G.N., J.N. de Klerk, M.B. Schneider, and J. van Eck, 2015: Detailed Assessment of the Biomass Resource and Potential Yield in a Selected Bush Encroached Area Of Namibia. Ministry of Agriculture, Water and Forestry, Windhoek, Namibia, 126 pp.
  1107. Stafford-Smith, M. et al., 2017: Integration: The Key to Implementing the Sustainable Development Goals. Sustain. Sci., 12, 911–919, doi:10.1007/s11625-016-0383-3.
  1108. Webb, N.P., C.J. Stokes, and N.A. Marshall, 2013: Integrating biophysical and socio-economic evaluations to improve the efficacy of adaptation assessments for agriculture. Glob. Environ. Chang., 23, 1164–1177, doi:10.1016/j.gloenvcha.2013.04.007.
  1109. Weltz, M., and K. Spaeth, 2012: Estimating effects of targeted conservation on nonfederal rangelands. Rangelands, 34, 35–40, doi:10.2111/RANGELANDS-D-12-00028.1.
  1110. Mainguet, M. and F. Dumay, 2011: Fighting Wind Erosion. One Aspect of the Combat Against Desertification. CSFD/Agropolis International, Les dossiers thématiques du CSFD/Agropolis International, Montpellier, France, 44 pp.
  1111. Aboushook, M., M.N. Al Maghrabi, A. Fadol, and G.S. Abdelhaffez, 2012: Different methods for stabilisation of sand dunes using calcium bentonite. Int. J. Environ. Eng., 4, 79, doi:10.1504/IJEE.2012.048103.
  1112. Rammal, M.M., and A.A. Jubair, 2015: Sand dunes stabilization using silica gel and cement kiln dust. Al-Nahrain J. Eng. Sci., 18, 179–191.
  1113. Sebaa A. et al., 2015: Guide des Techniques de Lutte Contre l’Ensablement au Sahara Algérien. CRSTRA, ISBN: 978-9931-438-05-2, Biskra, Algeria, 82 pp.
  1114. Yu Qiu, G., I.-B. Lee, H. Shimizu, Y. Gao, and G. Ding, 2004: Principles of sand dune fixation with straw checkerboard technology and its effects on the environment. J. Arid Environ., 56, 449–464, doi:10.1016/S0140-1963(03)00066-1.
  1115. Abdelkebir, T. and A. Ferchichi, 2015: Lutte biologique contre les accumulations sableuses dans la Tunisie aride, Revue des Régions Arides, Numéro Spécial 36: 210–220 (ISBN/ISNN 0330-7956).
  1116. Zhang, J. et al., 2015: Effects of sand dune stabilization on the spatial pattern of Artemisia ordosica population in Mu Us desert, Northwest China. PLoS One, 10, e0129728, doi:10.1371/journal.pone.0129728.
  1117. FAO, 2011: Gestion des Plantations sur Dunes. Document de Travail sur les Forêts et la Foresterie en Zones Arides. Document de travail sur les les Forêts et la Foresterie en zones arides, n° 3, Food and Agriculture Organization of the United Nations, Rome, Italy, 43 pp.
  1118. Yang, H. et al., 2014: Carbon sequestration capacity of shifting sand dune after establishing new vegetation in the Tengger Desert, northern China. Sci. Total Environ., 478, 1–11, doi:10.1016/J.SCITOTENV.2014.01.063.
  1119. Jan, S.U. et al., 2017: Analysis of salinity tolerance potential in synthetic hexaploid wheat. Pak. J. Bot, 49, 1269–1278.
  1120. Qadir, M., A. Ghafoor, and G. Murtaza, 2000: Amelioration strategies for saline soils: A review. L. Degrad. Dev., 11, 501–521, doi:10.1002/1099-145X(200011/12)11:6<501::AID-LDR405>3.0.CO;2-S.
  1121. Chughtai, M.I., K. Mahmood, and A.R. Awan, 2015: Growth performance of carp species fed on salt-tolerant roughages and formulated feed in brackish water under polyculture system. Pak. J. Zool., 47, 775–781.
  1122. Mahmood, R., R.A. Pielke, and C.A. McAlpine, 2016: Climate-relevant land use and land cover change policies. Bull. Am. Meteorol. Soc., 97, 195–202, doi:10.1175/BAMS-D-14-00221.1.
  1123. Sharma, R., S. Wungrampha, V. Singh, A. Pareek, and M.K. Sharma, 2016: Halophytes as bioenergy crops. Front. Plant Sci., 7, 1372, doi:10.3389/fpls.2016.01372.
  1124. Dagar, J.C., C.B. Pandey, and C.S. Chaturvedi, 2014: Agroforestry: A way forward for sustaining fragile coastal and island agroecosystems. In: Agroforestry Systems in India: Livelihood Security & Environmental Services Advances in Agroforestry [Dagar, J.C., A.K. Singh, and A. Arunachalam (eds.)]. Springer, New Delhi, India, pp. 185–232.
  1125. Wicke, B., E.M.W. Smeets, R. Akanda, L. Stille, R.K. Singh, A.R. Awan, 
K. Mahmood, and A.P.C. Faaij, 2013: Biomass production in agroforestry and forestry systems on salt-affected soils in South Asia: Exploration of the GHG balance and economic performance of three case studies. J. Environ. Manage., 127, 324–334, doi:10.1016/J.JENVMAN.2013.05.060.
  1126. Ashraf, M.Y., M. Ashraf, K. Mahmood, J. Akhter, F. Hussain, and M. Arshad, 2010: Phytoremediation of saline soils for sustainable agricultural productivity. In: Plant Adaptation and Phytoremediation [Ashraf M., M. Ozturk and M. Ahmad (eds.)]. Springer Netherlands, Dordrecht, Netherlands, pp. 335–355.
  1127. Awan, A.R., and K. Mahmood, 2017: Tree plantation in problem soils. In: Text Book of Applied Forestry [Tahir Siddiqui, M. and F. Nawaz (ed.)]. University of Agriculture, Faisalabad, Pakistan, pp. 140–159.
  1128. Mahmood, R., R.A. Pielke, and C.A. McAlpine, 2016: Climate-relevant land use and land cover change policies. Bull. Am. Meteorol. Soc., 97, 195–202, doi:10.1175/BAMS-D-14-00221.1.
  1129. Yaish, M.W., and P.P. Kumar, 2015: Salt tolerance research in date palm tree (Phoenix dactylifera L.), past, present, and future perspectives. Front. Plant Sci., 6, 348, doi:10.3389/fpls.2015.00348.
  1130. Al-Mulla, L., N.R. Bhat, and M. Khalil, 2013: Salt-tolerance of tissue-cultured date palm cultivars under controlled environment. World Acad. Sci. Eng. Technol. Int. J. Anim. Vet. Sci., 7, 811–814.
  1131. Alrasbi, S.A.R., N. Hussain, and H. Schmeisky, 2010: Evaluation of the growth of date palm seedlings irrigated with saline water in the Sultanate of Oman. IV International Date Palm Conference, 233–246, doi:10.17660/ActaHortic.2010.882.26.
  1132. Ibrahim, H.I.M., 2016: Tolerance of two pomegranates cultivars (Punica granatum L.) to salinity stress under hydroponic conditions. J. Basic Appl. Sci. Res, 6, 38–46. ISSN 2090-4304.
  1133. Okhovatian-Ardakani, A.R., M. Mehrabanian, F. Dehghani, and A. Akbarzadeh, 2010: Salt tolerance evaluation and relative comparison in cuttings of different pomegranate cultivars. Plant Soil Environ., 2010, 176–185. DOI: 10.17221/158/2009-PSE.
  1134. Abideen, Z. et al., 2014: Sustainable biofuel production from non-food sources? An overview. Emirates J. Food Agric., 26, 1057, http://www.ejfa.me/index.php/journal/article/view/617.
  1135. Dagar, J.C., C.B. Pandey, and C.S. Chaturvedi, 2014: Agroforestry: A way forward for sustaining fragile coastal and island agroecosystems. In: Agroforestry Systems in India: Livelihood Security & Environmental Services Advances in Agroforestry [Dagar, J.C., A.K. Singh, and A. Arunachalam (eds.)]. Springer, New Delhi, India, pp. 185–232.
  1136. Sanandiya, N.D., and A.K. Siddhanta, 2014: Chemical studies on the polysaccharides of Salicornia brachiata. Carbohydr. Polym., 112, 300–307, doi:10.1016/J.CARBPOL.2014.05.072.
  1137. Koyro, H.-W., T. Hussain, B. Huchzermeyer, and M.A. Khan, 2013: Photosynthetic and growth responses of a perennial halophytic grass Panicum turgidum to increasing NaCl concentrations. Environ. Exp. Bot., 91, 22–29, doi:10.1016/J.ENVEXPBOT.2013.02.007.
  1138. Akhter, J., K. Mahmood, K.A. Malik, S. Ahmed, and R. Murray, 2003: Amelioration of a saline sodic soil through cultivation of a salt-tolerant grass Leptochloa fusca. Environ. Conserv., 30, 168–174, doi:10.1017/S0376892903000158.
  1139. Sakai, Y. et al., 2012: Phytodesalination of a salt-affected soil with four halophytes in China. J. Arid L. Stud., 22, 239–302.
  1140. Wang, X., F. Yamauchi, J. Huang, and S. Rozelle, 2018a: What constrains mechanization in Chinese agriculture? Role of farm size and fragmentation. China Econ. Rev., in press, doi:10.1016/J.CHIECO.2018.09.002.
  1141. NIAB, 1997: Economic utilization of salt-affected soils In Twenty Five Years (1992–1997) of NIAB (Silver Jubilee Publication). Nuclear Institute for Agriculture and Biology, Faisalabad, Pakistan, pp. 123–151.
  1142. Akinshina, N., A. Azizov, T. Karasyova, and E. Klose, 2016: On the issue of halophytes as energy plants in saline environment. Biomass and Bioenergy, 91, 306–311, doi:10.1016/J.BIOMBIOE.2016.05.034.
  1143. Cordingley, J.E., K.A. Snyder, J. Rosendahl, F. Kizito, and D. Bossio, 2015: Thinking outside the plot: Addressing low adoption of sustainable land management in Sub-Saharan Africa. Curr. Opin. Environ. Sustain., 15, 35–40, doi:10.1016/J.COSUST.2015.07.010.
  1144. Fleskens, L., and L.C. Stringer, 2014: Land management and policy responses to mitigate desertification and land degradation. L. Degrad. Dev., 25, 1–4, doi:10.1002/ldr.2272.
  1145. Nyanga, A., A. Kessler, and A. Tenge, 2016: Key socio-economic factors influencing sustainable land management investments in the West Usambara Highlands, Tanzania. Land Use Policy, 51, 260–266, doi:10.1016/J.LANDUSEPOL.2015.11.020.
  1146. Barrett, C.B., L. Christiaensen, M. Sheahan, and A. Shimeles, 2017: On the structural transformation of rural Africa. J. Afr. Econ., 26, i11–i35, doi:10.1093/jae/ejx009.
  1147. Shiferaw, B., and A. Djido, 2016: Patterns of Labor Productivity and Income Diversification in the Rural Farm and Non-farm Sectors in Sub-Saharan Africa: Partnership for Economic Policy. Policy Brief No. 143. Nairobi, Kenya, 3 pp.
  1148. Altieri, M.A. and C.I. Nicholls, 2017: The adaptation and mitigation potential of traditional agriculture in a changing climate. Clim. Change, 140, 33–45, doi:10.1007/s10584-013-0909-y.
  1149. Engdawork, A., and H.-R. Bork, 2016: Farmers’ perception of land degradation and traditional knowledge in southern Ethiopia – Resilience and stability. 
L. Degrad. Dev., 27, 1552–1561, doi:10.1002/ldr.2364.
  1150. Belfer, E., J.D. Ford, and M. Maillet, 2017: Representation of indigenous peoples in climate change reporting. Clim. Change, 145, 57–70, doi:10.1007/s10584-017-2076-z.
  1151. Codjoe, S.N.A., G. Owusu, and V. Burkett, 2014: Perception, experience, and indigenous knowledge of climate change and variability: The case of Accra, a Sub-Saharan African city. Reg. Environ. Chang., 14, 369–383, doi:10.1007/s10113-013-0500-0.
  1152. Etchart, L., 2017: The role of indigenous peoples in combating climate change. Palgrave Commun., 3, 17085, doi:10.1057/palcomms.2017.85.
  1153. Ifejika Speranza, C., B. Kiteme, P. Ambenje, U. Wiesmann, and S. Makali, 2010: Indigenous knowledge related to climate variability and change: Insights from droughts in semi-arid areas of former Makueni District, Kenya. Clim. Change, 100, 295–315, doi:10.1007/s10584-009-9713-0.
  1154. Makondo, C.C., and D.S.G. Thomas, 2018: Climate change adaptation: Linking indigenous knowledge with Western science for effective adaptation. Environ. Sci. Policy, 88, 83–91, doi:10.1016/J.ENVSCI.2018.06.014.
  1155. Maldonado, J. et al., 2016: Engagement with indigenous peoples and honoring traditional knowledge systems. Clim. Change, 135, 111–126, doi:10.1007/s10584-015-1535-7.
  1156. Nyong, A., F. Adesina, and B. Osman Elasha, 2007: The value of indigenous knowledge in climate change mitigation and adaptation strategies in the African Sahel. Mitig. Adapt. Strateg. Glob. Chang., 12, 787–797, doi:10.1007/s11027-007-9099-0.
  1157. Altieri, M.A. and C.I. Nicholls, 2017: The adaptation and mitigation potential of traditional agriculture in a changing climate. Clim. Change, 140, 33–45, doi:10.1007/s10584-013-0909-y.
  1158. Biazin, B., G. Sterk, M. Temesgen, A. Abdulkedir, and L. Stroosnijder, 2012: Rainwater harvesting and management in rainfed agricultural systems in Sub-Saharan Africa – A review. Phys. Chem. Earth, Parts A/B/C, 47–48, 139–151, doi:10.1016/J.PCE.2011.08.015.
  1159. Enfors, E.I., and L.J. Gordon, 2008: Dealing with drought: The challenge of using water system technologies to break Dryland poverty traps. Glob. Environ. Chang., 18, 607–616, doi:10.1016/J.GLOENVCHA.2008.07.006.
  1160. Dominguez, P., 2014: Current situation and future patrimonializing perspectives for the governance of agropastoral resources in the Ait Ikis transhumants of the High Atlas (Morocco). In: Herrera P, Davies J, Baena P (eds.) The Governance of Rangelands. Collective Action for Sustainable Pastoralism. Routledge, Oxon and New York, pp. 126–144.
  1161. Auclair, L., P. Baudot, D. Genin, B. Romagny, and R. Simenel, 2011: Patrimony for resilience evidence from the Forest Agdal in the Moroccan High Atlas. Ecol. Soc., 16, doi:10.5751/ES-04429-160424.
  1162. Hussein, I.A.E., 2011: Desertification process in Egypt. In: Coping with Global Environmental Change, Disasters and Security: Threats, Challenges, Vulnerabilities and Risks [Brauch, H.G., U. Oswald Spring, C. Mesjasz, 
J. Grin, P. Kameri-Mbote, B. Chourou, P. Dunay and J. Brikmann (eds.)]. Springer, Berlin, Germany, pp. 863–874.
  1163. Louhaichi, M., and A. Tastad, 2010: The Syrian steppe: Past trends, current status, and future priorities. Rangelands, 32, 2–7, doi:10.2111/1551-501X-32.2.2.
  1164. Fre, Z., 2018: Knowledge Sovereignty among African Cattle Herders. UCL Press, London, UK, 200 pp.
  1165. Enfors, E.I., and L.J. Gordon, 2008: Dealing with drought: The challenge of using water system technologies to break Dryland poverty traps. Glob. Environ. Chang., 18, 607–616, doi:10.1016/J.GLOENVCHA.2008.07.006.
  1166. Engdawork, A., and H.-R. Bork, 2016: Farmers’ perception of land degradation and traditional knowledge in southern Ethiopia – Resilience and stability. 
L. Degrad. Dev., 27, 1552–1561, doi:10.1002/ldr.2364.
  1167. Dominguez, P., 2014: Current situation and future patrimonializing perspectives for the governance of agropastoral resources in the Ait Ikis transhumants of the High Atlas (Morocco). In: Herrera P, Davies J, Baena P (eds.) The Governance of Rangelands. Collective Action for Sustainable Pastoralism. Routledge, Oxon and New York, pp. 126–144.
  1168. Fernández-Giménez, M.E., and F. Fillat Estaque, 2012: Pyrenean Pastoralists’ Ecological knowledge: Documentation and application to natural resource management and adaptation. Hum. Ecol., 40, 287–300, doi:10.1007/s10745-012-9463-x.
  1169. Hussein, I.A.E., 2011: Desertification process in Egypt. In: Coping with Global Environmental Change, Disasters and Security: Threats, Challenges, Vulnerabilities and Risks [Brauch, H.G., U. Oswald Spring, C. Mesjasz, 
J. Grin, P. Kameri-Mbote, B. Chourou, P. Dunay and J. Brikmann (eds.)]. Springer, Berlin, Germany, pp. 863–874.
  1170. Kodirekkala, K.R., 2017: Internal and external factors affecting loss of traditional knowledge: Evidence from a Horticultural Society in South India. J. Anthropol. Res., 73, 22–42, doi:10.1086/690524.
  1171. Moreno-Calles, A.I., A. Casas, E. García-Frapolli, and I. Torres-García, 2012: Traditional agroforestry systems of multi-crop ‘milpa’ and ‘chichipera’ cactus forest in the arid Tehuacán Valley, Mexico: Their management and role in people’s subsistence. Agrofor. Syst., 84, 207–226, doi:10.1007/s10457-011-9460-x.
  1172. Engdawork, A., and H.-R. Bork, 2016: Farmers’ perception of land degradation and traditional knowledge in southern Ethiopia – Resilience and stability. 
L. Degrad. Dev., 27, 1552–1561, doi:10.1002/ldr.2364.
  1173. Guzman, C.D. et al., 2018: Developing soil conservation strategies with technical and community knowledge in a degrading sub-humid mountainous landscape. L. Degrad. Dev., 29, 749–764, doi:10.1002/ldr.2733.
  1174. Adger, W.N., 2003: Social capital, collective action, and adaptation to climate change. Econ. Geogr., 79, 387–404, doi:10.1111/j.1944-8287.2003.tb00220.x.
  1175. Engdawork, A., and H.-R. Bork, 2016: Farmers’ perception of land degradation and traditional knowledge in southern Ethiopia – Resilience and stability. 
L. Degrad. Dev., 27, 1552–1561, doi:10.1002/ldr.2364.
  1176. Eriksen, S., and J. Lind, 2009: Adaptation as a political process: Adjusting to drought and conflict in Kenya’s drylands. Environ. Manage., 43, 817–835, doi:10.1007/s00267-008-9189-0.
  1177. Ostrom, E., 2009: A general framework for analyzing sustainability of social-ecological systems. Science, 325, 419–422, doi:10.1126/science.1172133.
  1178. Rodima-Taylor, D., M.F. Olwig, and N. Chhetri, 2012: Adaptation as innovation, innovation as adaptation: An institutional approach to climate change. Appl. Geogr., 33, 107–111, doi:10.1016/J.APGEOG.2011.10.011.
  1179. van Rijn, F., E. Bulte, and A. Adekunle, 2012: Social capital and agricultural innovation in Sub-Saharan Africa. Agric. Syst., 108, 112–122, doi:10.1016/J.AGSY.2011.12.003.
  1180. Woolcock, M., and D. Narayan, 2000: Social capital: Implications for development theory, research, and policy. World Bank Res. Obs., 15, 25–249.
  1181. Ahlerup, P., O. Olsson, and D. Yanagizawa, 2009: Social capital vs institutions in the growth process. Eur. J. Polit. Econ., 25, 1–14, doi:10.1016/J.EJPOLECO.2008.09.008.
  1182. Ajayi, O.C., F.K. Akinnifesi, and A.O. Ajayi, 2016: How by-laws and collective action influence farmers’ adoption of agroforestry and natural resource management technologies: Lessons from Zambia. For. Trees Livelihoods, 25, 102–113, doi:10.1080/14728028.2016.1153435.
  1183. Infante, F., 2017: The role of social capital and labour exchange in the soils of Mediterranean Chile. Rural Soc., 26, 107–124, doi:10.1080/10371656.2017.1330837.
  1184. Kassie, M., M. Jaleta, B. Shiferaw, F. Mmbando, and M. Mekuria, 2013: Adoption of interrelated sustainable agricultural practices in smallholder systems: Evidence from rural Tanzania. Technol. Forecast. Soc. Change, 80, 525–540, doi:10.1016/J.TECHFORE.2012.08.007.
  1185. Nyangena, W., 2008: Social determinants of soil and water conservation in rural Kenya. Environ. Dev. Sustain., 10, 745–767, doi:10.1007/s10668-007-9083-6.
  1186. Willy, D.K., and K. Holm-Müller, 2013: Social influence and collective action effects on farm level soil conservation effort in rural Kenya. Ecol. Econ., 90, 94–103, doi:10.1016/J.ECOLECON.2013.03.008.
  1187. Wossen, T., T. Berger, and S. Di Falco, 2015: Social capital, risk preference and adoption of improved farm land management practices in Ethiopia. Agric. Econ., 46, 81–97, doi:10.1111/agec.12142.
  1188. Teshome, A., J. de Graaff, and A. Kessler, 2016: Investments in land management in the north-western highlands of Ethiopia: The role of social capital. Land Use Policy, 57, 215–228, doi:10.1016/J.LANDUSEPOL.2016.05.019.
  1189. Kihiu, E., 2016a: Pastoral Practices, Economics, and Institutions of Sustainable Rangeland Management in Kenya. PhD Thesis, University of Bonn, Bonn, Germany, 167 pp.
  1190. Stringer, L.C., C. Twyman, and D.S.G. Thomas, 2007: Combating land degradation through participatory means: The case of Swaziland. Ambio, 36, 387–393. doi.org/10.1579/0044-7447(2007)36[387:CLDTPM]2.0.CO;2.
  1191. van Rijn, F., E. Bulte, and A. Adekunle, 2012: Social capital and agricultural innovation in Sub-Saharan Africa. Agric. Syst., 108, 112–122, doi:10.1016/J.AGSY.2011.12.003.
  1192. Ulambayar, T., M.E. Fernández-Giménez, B. Baival, and B. Batjav, 2017: Social outcomes of community-based rangeland management in Mongolian steppe ecosystems. Conserv. Lett., 10, 317–327, doi:10.1111/conl.12267.
  1193. Reij, C., and A. Waters-Bayer (eds.), 2001: Farmer Innovation in Africa: A Source of Inspiration for Agricultural Development. Earthscan Publications Ltd, London, UK. ISBN 9781853838163, 384 pp.
  1194. Tambo, J.A., and T. Wünscher, 2015: Identification and prioritization of farmers’ innovations in northern Ghana. Renew. Agric. Food Syst., 30, 537–549, doi:10.1017/S1742170514000374.
  1195. Waters-Bayer, A., L. van Veldhuizen, M. Wongtschowski, and C. Wettasinha, 2009: Recognizing and enhancing processes of local innovation. In: Innovation Africa: Enriching Farmers’ Livelihoods [Sanginga, P., A. Waters-Bayer, S. Kaaria, J. Njuki, and C. Wettasinha (eds.)]. Routledge, London, UK, pp. 239–254.
  1196. Bonney, R., T.B. Phillips, H.L. Ballard, and J.W. Enck, 2016: Can citizen science enhance public understanding of science? Public Underst. Sci., 25, 2–16, doi:10.1177/0963662515607406.
  1197. de Vente, J., M.S. Reed, L.C. Stringer, S. Valente, and J. Newig, 2016: How does the context and design of participatory decision-making processes affect their outcomes? Evidence from sustainable land management in global drylands. Ecol. Soc., 21, 1–24, doi:10.5751/ES-08053-210224.
  1198. Gupta, A.K. et al., 2016: Theory of open inclusive innovation for reciprocal, responsive and respectful outcomes: Coping creatively with climatic and institutional risks. J. Open Innov. Technol. Mark. Complex., 2, 16, doi:10.1186/s40852-016-0038-8.
  1199. Dey, A., A. Gupta, and G. Singh, 2017: Open innovation at different levels for higher climate risk resilience. Sci. Technol. Soc., 22, 388–406, doi:10.1177/0971721817723242.
  1200. Cornell, S. et al., 2013: Opening up knowledge systems for better responses to global environmental change. Environ. Sci. Policy, 28, 60–70, doi:10.1016/J.ENVSCI.2012.11.008.
  1201. Herrick, J.E. et al., 2017: Two new mobile apps for rangeland inventory and monitoring by landowners and land managers. Rangelands, 39, 46–55, doi:10.1016/j.rala.2016.12.003.
  1202. McKinley, D.C. et al., 2017: Citizen science can improve conservation science, natural resource management, and environmental protection. Biol. Conserv., 208, 15–28, doi:10.1016/J.BIOCON.2016.05.015.
  1203. Steger, C., B. Butt, and M.B. Hooten, 2017: Safari science: Assessing the reliability of citizen science data for wildlife surveys. J. Appl. Ecol., 54, 2053–2062, doi:10.1111/1365-2664.12921.
  1204. Banadda, N., 2010: Gaps, barriers and bottlenecks to sustainable land management (SLM) adoption in Uganda. African J. Agric. Res., 5, 3571–3580, doi:10.5897/ajar10.029.
  1205. Cordingley, J.E., K.A. Snyder, J. Rosendahl, F. Kizito, and D. Bossio, 2015: Thinking outside the plot: Addressing low adoption of sustainable land management in Sub-Saharan Africa. Curr. Opin. Environ. Sustain., 15, 35–40, doi:10.1016/J.COSUST.2015.07.010.
  1206. Lokonon, B.O.K., and A.A. Mbaye, 2018: Climate change and adoption of sustainable land management practices in the Niger Basin of Benin. Nat. Resour. Forum, 42, 42–53, doi:10.1111/1477-8947.12142.
  1207. Mulinge, W., P. Gicheru, F. Murithi, P. Maingi, E. Kihiu, O.K. Kirui, and A. Mirzabaev, 2016: Economics of land degradation and improvement in Kenya. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 471–498.
  1208. Wildemeersch, J.C.J. et al., 2015: Assessing the constraints to adopt water and soil conservation techniques in Tillaberi, Niger. L. Degrad. Dev., 26, 491–501, doi:10.1002/ldr.2252.
  1209. Boserup, E., 1965: The Conditions of Agricultural Growth: The Economics of Agrarian Change Under Population Pressure. George All & Unwin, Ltd., London, UK, 124 pp.
  1210. Safriel, U., and Z. Adeel, 2008: Development paths of drylands: Thresholds and sustainability. Sustain. Sci., 3, 117–123, doi:10.1007/s11625-007-0038-5.
  1211. Belay, A., J.W. Recha, T. Woldeamanuel, and J.F. Morton, 2017: Smallholder farmers’ adaptation to climate change and determinants of their adaptation decisions in the Central Rift Valley of Ethiopia. Agric. Food Secur., 6, 24, doi:10.1186/s40066-017-0100-1.
  1212. Bryan, E., T.T. Deressa, G.A. Gbetibouo and C. Ringler, 2009: Adaptation to climate change in Ethiopia and South Africa: Options and constraints. Environ. Sci. Policy, 12, 413–426, doi:10.1016/j.envsci.2008.11.002.
  1213. Dumenu, W.K. and E.A. Obeng, 2016: Climate change and rural communities in Ghana: Social vulnerability, impacts, adaptations and policy implications. Environ. Sci. Policy, 55, 208–217, doi:10.1016/J.ENVSCI.2015.10.010.
  1214. Salik, K.M., A. Qaisrani, M.A. Umar, and S.M. Ali, 2017: Migration Futures in Asia and Africa: Economic Opportunities and Distributional Effects – the Case of Pakistan. Working Paper, Sustainable Development Policy Institute, Islamabad, Pakistan, 59 pp.
  1215. Shiferaw, B.A., J. Okello, and R.V. Reddy, 2009: Adoption and adaptation of natural resource management innovations in smallholder agriculture: Reflections on key lessons and best practices. Environ. Dev. Sustain., 11, 601–619, doi:10.1007/s10668-007-9132-1.
  1216. Fratkin, E., 2013: Seeking alternative livelihoods in pastoral areas. In: Pastoralism and Development in Africa Dynamic Change at the Margins [Catley, A., J. Lind, and I. Scoones (eds.)]. Routledge, London, UK, 328 pp.
  1217. Lybbert, T.J., C.B. Barrett, S. Desta, and D. Layne Coppock, 2004: Stochastic wealth dynamics and risk management among a poor population. Econ. J., 114, 750–777, doi:10.1111/j.1468-0297.2004.00242.x.
  1218. Reardon, T., J.E. Taylor, K. Stamoulis, P. Lanjouw, and A. Balisacan, 2008: Effects of non-farm employment on rural income inequality in developing countries: An investment perspective. J. Agric. Econ., 51, 266–288, doi:10.1111/j.1477-9552.2000.tb01228.x.
  1219. Walther, G.-R. et al., 2002: Ecological responses to recent climate change. Nature, 416, 389–395, doi:10.1038/416389a.
  1220. Hornbeck, R., 2012: The enduring impact of the American Dust Bowl: Short- and long-run adjustments to environmental catastrophe. Am. Econ. Rev., 102, 1477–1507, doi:10.1257/aer.102.4.1477.
  1221. Barbier, B., H. Yacouba, H. Karambiri, M. Zoromé, and B. Somé, 2009: Human vulnerability to climate variability in the Sahel: Farmers’ adaptation strategies in northern Burkina Faso. Environ. Manage., 43, 790–803, doi:10.1007/s00267-008-9237-9.
  1222. Nawrotzki, R.J., and J. DeWaard, 2016: Climate shocks and the timing of migration from Mexico. Popul. Environ., 38, 72–100, doi:10.1007/s11111-016-0255-x.
  1223. Lima, M., D.A. Christie, M.C. Santoro, and C. Latorre, 2016: Coupled socio-environmental changes triggered indigenous Aymara depopulation of the semi-arid Andes of Tarapacá-Chile during the late 19th–20th centuries. PLoS One, 11, e0160580, doi:10.1371/journal.pone.0160580.
  1224. Liehr, S., L. Drees, and D. Hummel, 2016: Migration as societal response to climate change and land degradation in Mali and Senegal. In: Adaptation to Climate Change and Variability in Rural West Africa [Yaro, J.A. and 
J. Hesselberg (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 147–169.
  1225. Barrios, S., L. Bertinelli, and E. Strobl, 2006: Climatic change and rural–urban migration: The case of Sub-Saharan Africa. J. Urban Econ., 60, 357–371, doi:10.1016/J.JUE.2006.04.005.
  1226. Henderson, J.V., A. Storeygard, and U. Deichmann, 2017: Has climate change driven urbanisation in Africa? J. Dev. Econ., 124, 60–82, doi:10.1016/J.JDEVECO.2016.09.001.
  1227. Madani, K., A. AghaKouchak, and A. Mirchi, 2016: Iran’s socio-economic drought: Challenges of a water-bankrupt nation. Iran. Stud., 49, 997–1016, doi:10.1080/00210862.2016.1259286.
  1228. Cattaneo, C., and G. Peri, 2016: The migration response to increasing temperatures. J. Dev. Econ., 122, 127–146, doi:10.1016/J.JDEVECO.2016.05.004.
  1229. Liehr, S., L. Drees, and D. Hummel, 2016: Migration as societal response to climate change and land degradation in Mali and Senegal. In: Adaptation to Climate Change and Variability in Rural West Africa [Yaro, J.A. and 
J. Hesselberg (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 147–169.
  1230. Salik, K.M., A. Qaisrani, M.A. Umar, and S.M. Ali, 2017: Migration Futures in Asia and Africa: Economic Opportunities and Distributional Effects – the Case of Pakistan. Working Paper, Sustainable Development Policy Institute, Islamabad, Pakistan, 59 pp.
  1231. Sikder, M.J.U., and V. Higgins, 2017: Remittances and social resilience of migrant households in rural Bangladesh. Migr. Dev., 6, 253–275, doi:10.1080/21632324.2016.1142752.
  1232. Chen, R., C. Ye, Y. Cai, X. Xing, and Q. Chen, 2014: The impact of rural out-migration on land use transition in China: Past, present and trend. Land Use Policy, 40, 101–110, doi:10.1016/J.LANDUSEPOL.2013.10.003.
  1233. Sanz, M.J. et al., 2017: Sustainable Land Management Contribution to Successful Land-Based Climate Change Adaptation and Mitigation. A Report of the Science-Policy Interface. United Nations Convention to Combat Desertification, Bonn, Germany, 170 pp.
  1234. Nkonya, E. et al., 2016a: Global cost of land degradation. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 117–165.
  1235. Mirzabaev, A., M. Ahmed, J. Werner, J. Pender, and M. Louhaichi, 2016a: Rangelands of Central Asia: Challenges and opportunities. J. Arid Land, 8, 93–108, doi:10.1007/s40333-015-0057-5.
  1236. Gebreselassie, S., O.K. Kirui, and A. Mirzabaev, 2016: Economics of land degradation and improvement in Ethiopia. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 401–430.
  1237. Mythili, G., and J. Goedecke, 2016: Economics of Land Degradation in India. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 431–469.
  1238. Mulinge, W., P. Gicheru, F. Murithi, P. Maingi, E. Kihiu, O.K. Kirui, and A. Mirzabaev, 2016: Economics of land degradation and improvement in Kenya. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 471–498.
  1239. Moussa, B., E. Nkonya, S. Meyer, E. Kato, T. Johnson, and J. Hawkins, 2016: Economics of land degradation and improvement in Niger. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 499–539.
  1240. Sow, S., E. Nkonya, S. Meyer, and E. Kato, 2016: Cost, drivers and action against land degradation in Senegal. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 577–608.
  1241. Cordingley, J.E., K.A. Snyder, J. Rosendahl, F. Kizito, and D. Bossio, 2015: Thinking outside the plot: Addressing low adoption of sustainable land management in Sub-Saharan Africa. Curr. Opin. Environ. Sustain., 15, 35–40, doi:10.1016/J.COSUST.2015.07.010.
  1242. Giger, M., H. Liniger, C. Sauter, and G. Schwilch, 2015: Economic benefits and costs of sustainable land management technologies: An analysis of WOCAT’s global data. L. Degrad. Dev., 29, 962–974, doi:10.1002/ldr.2429.
  1243. Lokonon, B.O.K., and A.A. Mbaye, 2018: Climate change and adoption of sustainable land management practices in the Niger Basin of Benin. Nat. Resour. Forum, 42, 42–53, doi:10.1111/1477-8947.12142.
  1244. Nkonya, E. et al., 2016a: Global cost of land degradation. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 117–165.
  1245. Schmidt, E., P. Chinowsky, S. Robinson, and K. Strzepek, 2017: Determinants and impact of sustainable land management (SLM) investments: A systems evaluation in the Blue Nile Basin, Ethiopia. Agric. Econ., 48, 613–627, doi:10.1111/agec.12361.
  1246. Giger, M., H. Liniger, C. Sauter, and G. Schwilch, 2015: Economic benefits and costs of sustainable land management technologies: An analysis of WOCAT’s global data. L. Degrad. Dev., 29, 962–974, doi:10.1002/ldr.2429.
  1247. Sanz, M.J. et al., 2017: Sustainable Land Management Contribution to Successful Land-Based Climate Change Adaptation and Mitigation. A Report of the Science-Policy Interface. United Nations Convention to Combat Desertification, Bonn, Germany, 170 pp.
  1248. Schmidt, E., P. Chinowsky, S. Robinson, and K. Strzepek, 2017: Determinants and impact of sustainable land management (SLM) investments: A systems evaluation in the Blue Nile Basin, Ethiopia. Agric. Econ., 48, 613–627, doi:10.1111/agec.12361.
  1249. Djanibekov, U., and A. Khamzina, 2016: Stochastic economic assessment of afforestation on marginal land in irrigated farming system. Environ. Resour. Econ., 63, 95–117, doi:10.1007/s10640-014-9843-3.
  1250. Djanibekov, U. et al., 2016: Adoption of sustainable land uses in post-Soviet Central Asia: The case for agroforestry. Sustainability, 8, 1030, doi:10.3390/su8101030.
  1251. Giger, M., H. Liniger, C. Sauter, and G. Schwilch, 2015: Economic benefits and costs of sustainable land management technologies: An analysis of WOCAT’s global data. L. Degrad. Dev., 29, 962–974, doi:10.1002/ldr.2429.
  1252. Moussa, B., E. Nkonya, S. Meyer, E. Kato, T. Johnson, and J. Hawkins, 2016: Economics of land degradation and improvement in Niger. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 499–539.
  1253. Sow, S., E. Nkonya, S. Meyer, and E. Kato, 2016: Cost, drivers and action against land degradation in Senegal. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 577–608.
  1254. Giger, M., H. Liniger, C. Sauter, and G. Schwilch, 2015: Economic benefits and costs of sustainable land management technologies: An analysis of WOCAT’s global data. L. Degrad. Dev., 29, 962–974, doi:10.1002/ldr.2429.
  1255. Nkonya, E. et al., 2016a: Global cost of land degradation. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 117–165.
  1256. Moussa, B., E. Nkonya, S. Meyer, E. Kato, T. Johnson, and J. Hawkins, 2016: Economics of land degradation and improvement in Niger. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 499–539.
  1257. Bouma, J., and J.H.M. Wösten, 2016: How to characterize ‘good’ and ‘greening’ in the EU Common Agricultural Policy (CAP): The case of clay soils in the Netherlands. Soil Use Manag., 32, 546–552, doi:10.1111/sum.12289.
  1258. Lambin, E.F. et al., 2014: Effectiveness and synergies of policy instruments for land use governance in tropical regions. Glob. Environ. Chang., 28, 129–140, doi:10.1016/J.GLOENVCHA.2014.06.007.
  1259. Reed, M.S., 2015: Reorienting land degradation towards sustainable land management: Linking sustainable livelihoods with ecosystem services in rangeland systems. J. Environ. Manage., 151, 472–485, doi:10.1016/J.JENVMAN.2014.11.010.
  1260. Schiappacasse, I., L. Nahuelhual, F. Vásquez, and C. Echeverría, 2012: Assessing the benefits and costs of Dryland forest restoration in central Chile. J. Environ. Manage., 97, 38–45. https://doi.org/10.1016/j.jenvman.2011.11.007.
  1261. van Zanten, B.T. et al., 2014: European agricultural landscapes, common agricultural policy and ecosystem services: A review. Agron. Sustain. Dev., 34, 309–325, doi:10.1007/s13593-013-0183-4.
  1262. Cowie, A.L. et al., 2018: Land in balance: The scientific conceptual framework for. Environ. Sci. Policy, 79, 25–35, doi:10.1016/J.ENVSCI.2017.10.011.
  1263. Orr, B.J. et al., 2017: Scientific Conceptual Framework for Land Degradation Neutrality. A Report of the Science-Policy Interface. United Nations Convention to Combat Desertification, Bonn, Germany, 128 pp.
  1264. Chasek, P., M. Akhtar-Schuster, B.J. Orr, A. Luise, H. Rakoto Ratsimba, and U. Safriel, 2019: Land degradation neutrality: The science-policy interface from the UNCCD to national implementation. Environ. Sci. Policy, 92, 182–190, doi:10.1016/j.envsci.2018.11.017.
  1265. Marques, M. et al., 2016: Multifaceted impacts of sustainable land management in drylands: A review. Sustainability, 8, 177, doi:10.3390/su8020177.
  1266. Aw-Hassan, A. et al., 2016: Economics of land degradation and improvement in Uzbekistan. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 651–682.
  1267. Gebreselassie, S., O.K. Kirui, and A. Mirzabaev, 2016: Economics of land degradation and improvement in Ethiopia. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 401–430.
  1268. Mythili, G., and J. Goedecke, 2016: Economics of Land Degradation in India. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 431–469.
  1269. Nkonya, E., and W. Anderson, 2015: Exploiting provisions of land economic productivity without degrading its natural capital. J. Arid Environ., 112, 33–43, doi:10.1016/J.JARIDENV.2014.05.012.
  1270. Nkonya, E., F. Place, E. Kato, and M. Mwanjololo, 2015: Climate risk management through sustainable land management in Sub-Saharan Africa. In: Sustainable Intensification to Advance Food Security and Enhance Climate Resilience in Africa [Lal, R., B.R. Singh, D.L. Mwaseba, 
D. Kraybill, D.O. Hansen and L.O. Eik (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 75–111.
  1271. Hornbeck, R., 2012: The enduring impact of the American Dust Bowl: Short- and long-run adjustments to environmental catastrophe. Am. Econ. Rev., 102, 1477–1507, doi:10.1257/aer.102.4.1477.
  1272. Little, P., 2010: Unofficial cross-border trade in eastern Africa. In: Food Security in Africa: Market and Trade Policy for Staple Foods in Eastern and Southern Africa [Sarris, A., T. Jayne, and J. Morrison, (eds.)]. Edwin Elgar Publishing, Cheltenham, UK, pp. 158–181.
  1273. Ait Hou, M., C. Grazia, and G. Malorgio, 2015: Food safety standards and international supply chain organization: A case study of the Moroccan fruit and vegetable exports. Food Control, 55, 190–199, doi:10.1016/J.FOODCONT.2015.02.023.
  1274. Keiichiro, H., T. Otsuki, and J.S. Wilson, 2015: Food safety standards and international trade: The impact on developing countries’ export performance. In: Food Safety, Market Organization, Trade and Development [Hammoudi, A., C, Grazia, Y. Surry, J.B. Traversac (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 151–166.
  1275. Unnevehr, L., 2015: Food safety in developing countries: Moving beyond exports. Glob. Food Sec., 4, 24–29, doi:10.1016/J.GFS.2014.12.001.
  1276. Bobojonov, I., R. Teuber, S. Hasanov, V. Urutyan, and T. Glauben, 2016: Farmers’ export market participation decisions in transition economies: A comparative study between Armenia and Uzbekistan. Dev. Stud. Res., 3, 25–35, doi:10.1080/21665095.2016.1262272.
  1277. Christy, R.D., C.A. da Silva, N. Mhlanga, E. Mabaya, and K. Tihanyi (eds.), 2014: Innovative Institutions, Public Policies and Private Strategies for Agro-Enterprise Development. Co-published with Food and Agriculture Organization of the United Nations, Rome, Italy, 368 pp. doi:10.1142/9131.
  1278. Nakasone, E., M. Torero, and B. Minten, 2014: The power of information: The ICT revolution in agricultural development. Annu. Rev. Resour. Econ., 6, 533–550, doi:10.1146/annurev-resource-100913-012714.
  1279. Broeckhoven, N., and A. Cliquet, 2015: Gender and ecological restoration: Time to connect the dots. Restor. Ecol., 23, 729–736, doi:10.1111/rec.12270.
  1280. Carr, E.R., and M.C. Thompson, 2014: Gender and climate change adaptation in agrarian settings: Current thinking, new directions, and research frontiers. Geogr. Compass, 8, 182–197, doi:10.1111/gec3.12121.
  1281. Catacutan, D.C., and G.B. Villamor, 2016: Gender roles and land use preferences – Implications to landscape restoration in Southeast Asia. In: Land Restoration: Reclaiming Landscapes for a Sustainable Future. [Chabay, I., M. Frick and J. Helgeson (eds.)]. Elsevier, Oxford, UK, pp. 431–440.
  1282. Dah-gbeto, A.P., and G.B. Villamor, 2016: Gender-specific responses to climate variability in a semi-arid ecosystem in northern Benin. Ambio, 45, 297–308, doi:10.1007/s13280-016-0830-5.
  1283. Nelson, V., L. Forsythe, and J. Morton, 2015: Empowering Dryland Women: Capturing Opportunities in Land Rights, Governance and Resilience. A Synthesis of Thematic Papers from the Series ‘Women’s Empowerment in the Drylands. Natural Resource Institute, University of Greenwich, Chatham, UK, 11 pp.
  1284. Thompson-Hall, M., E.R. Carr, and U. Pascual, 2016: Enhancing and expanding intersectional research for climate change adaptation in agrarian settings. Ambio, 45, 373–382, doi:10.1007/s13280-016-0827-0.
  1285. Villamor, G.B., and M. van Noordwijk, 2016: Gender specific land-use decisions and implications for ecosystem services in semi-matrilineal Sumatra. Glob. Environ. Chang., 39, 69–80, doi:10.1016/J.GLOENVCHA.2016.04.007.
  1286. Dah-gbeto, A.P., and G.B. Villamor, 2016: Gender-specific responses to climate variability in a semi-arid ecosystem in northern Benin. Ambio, 45, 297–308, doi:10.1007/s13280-016-0830-5.
  1287. Manlosa, A.O., J. Schultner, I. Dorresteijn, and J. Fischer, 2018: Leverage points for improving gender equality and human well-being in a smallholder farming context. Sustain. Sci., 1–13, doi:10.1007/s11625-018-0636-4.
  1288. Kassie, M., H. Teklewold, M. Jaleta, P. Marenya, and O. Erenstein, 2015: Understanding the adoption of a portfolio of sustainable intensification practices in eastern and southern Africa. Land Use Policy, 42, 400–411, doi:10.1016/J.LANDUSEPOL.2014.08.016.
  1289. Nkonya, E., and W. Anderson, 2015: Exploiting provisions of land economic productivity without degrading its natural capital. J. Arid Environ., 112, 33–43, doi:10.1016/J.JARIDENV.2014.05.012.
  1290. Nkonya, E., F. Place, E. Kato, and M. Mwanjololo, 2015: Climate risk management through sustainable land management in Sub-Saharan Africa. In: Sustainable Intensification to Advance Food Security and Enhance Climate Resilience in Africa [Lal, R., B.R. Singh, D.L. Mwaseba, 
D. Kraybill, D.O. Hansen and L.O. Eik (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 75–111.
  1291. Nyanga, A., A. Kessler, and A. Tenge, 2016: Key socio-economic factors influencing sustainable land management investments in the West Usambara Highlands, Tanzania. Land Use Policy, 51, 260–266, doi:10.1016/J.LANDUSEPOL.2015.11.020.
  1292. Berman, R., C. Quinn, and J. Paavola, 2012: The role of institutions in the transformation of coping capacity to sustainable adaptive capacity. Environ. Dev., 2, 86–100, doi:10.1016/J.ENVDEV.2012.03.017.
  1293. Chen, J., S. Yin, H. Gebhardt, and X. Yang, 2018: Farmers’ livelihood adaptation to environmental change in an arid region: A case study of the Minqin Oasis, north-western China. Ecol. Indic., 93, 411–423, doi:10.1016/J.ECOLIND.2018.05.017.
  1294. Chen, T. et al., 2019: Disentangling the relative impacts of climate change and human activities on arid and semi-arid grasslands in Central Asia during 1982–2015. Sci. Total Environ., 653, 1311–1325, doi:10.1016/j.scitotenv.2018.11.058.
  1295. Descheemaeker, K., M. Zijlstra, P. Masikati, O. Crespo, and S. Homann-Kee Tui, 2018: Effects of climate change and adaptation on the livestock component of mixed farming systems: A modelling study from semi-arid Zimbabwe. Agric. Syst., 159, 282–295, doi:10.1016/J.AGSY.2017.05.004.
  1296. Popp, A., S. Domptail, N. Blaum, and F. Jeltsch, 2009: Landuse experience does qualify for adaptation to climate change. Ecol. Modell., 220, 694–702, doi:10.1016/J.ECOLMODEL.2008.11.015.
  1297. Tambo, J.A., 2016: Adaptation and resilience to climate change and variability in north-east Ghana. Int. J. Disaster Risk Reduct., 17, 85–94, doi:10.1016/J.IJDRR.2016.04.005.
  1298. Yaro, J.A., J. Teye, and S. Bawakyillenuo, 2015: Local institutions and adaptive capacity to climate change/variability in the northern savanna of Ghana. Clim. Dev., 7, 235–245, doi:10.1080/17565529.2014.951018.
  1299. Osbahr, H., C. Twyman, W. Neil Adger, and D.S.G. Thomas, 2008: Effective livelihood adaptation to climate change disturbance: Scale dimensions of practice in Mozambique. Geoforum, 39, 1951–1964, doi:10.1016/J.GEOFORUM.2008.07.010.
  1300. Peters, D.P.C., K.M. Havstad, S.R. Archer, and O.E. Sala, 2015: Beyond desertification: New paradigms for Dryland landscapes. Front. Ecol. Environ., 13, 4–12, doi:10.1890/140276.
  1301. Tenge, A.J., J. De Graaff, and J.P. Hella, 2004: Social and economic factors affecting the adoption of soil and water conservation in West Usambara highlands, Tanzania. L. Degrad. Dev., 15, 99–114, doi:10.1002/ldr.606.
  1302. Morton, J., 2017: Climate change and African agriculture. In: Making Climate Compatible Development Happen [Nunan, F. (ed.)]. Routledge, London, UK, pp. 87–113.
  1303. Reichardt, K., 2010: College on soil physics: Soil physical properties and processes under climate change. Soil and Tillage Research, 79, 131–143, doi:10.1016/j.still.2004.07.002.
  1304. Mapfumo, P., F. Mtambanengwe, and R. Chikowo, 2016: Building on indigenous knowledge to strengthen the capacity of smallholder farming communities to adapt to climate change and variability in southern Africa. Clim. Dev., 8, 72–82, doi:10.1080/17565529.2014.998604.
  1305. Bonney, R., T.B. Phillips, H.L. Ballard, and J.W. Enck, 2016: Can citizen science enhance public understanding of science? Public Underst. Sci., 25, 2–16, doi:10.1177/0963662515607406.
  1306. Rusike, J., S. Twomlow, H.A. Freeman, and G.M. Heinrich, 2006: Does farmer participatory research matter for improved soil fertility technology development and dissemination in southern Africa? Int. J. Agric. Sustain., 4, 176–192, doi:10.1080/14735903.2006.9684801.
  1307. Morton, J., 2017: Climate change and African agriculture. In: Making Climate Compatible Development Happen [Nunan, F. (ed.)]. Routledge, London, UK, pp. 87–113.
  1308. Bestelmeyer, B.T. et al., 2017: State and Transition Models: Theory, Applications, and Challenges. Springer International Publishing, Cham, Switzerland, pp. 303–345.
  1309. Bambio, Y., and S. Bouayad Agha, 2018: Land tenure security and investment: Does strength of land right really matter in rural Burkina Faso? World Dev., 111, 130–147, doi:10.1016/J.WORLDDEV.2018.06.026.
  1310. Higgins, D., T. Balint, H. Liversage, and P. Winters, 2018: Investigating the impacts of increased rural land tenure security: A systematic review of the evidence. J. Rural Stud., 61, 34–62, doi:10.1016/J.JRURSTUD.2018.05.001.
  1311. Holden, S.T., and H. Ghebru, 2016: Land tenure reforms, tenure security and food security in poor agrarian economies: Causal linkages and research gaps. Glob. Food Sec., 10, 21–28, doi:10.1016/J.GFS.2016.07.002.
  1312. Paltasingh, K.R., 2018: Land tenure security and adoption of modern rice technology in Odisha, Eastern India: Revisiting Besley’s hypothesis. Land Use Policy, 78, 236–244, doi:10.1016/J.LANDUSEPOL.2018.06.031.
  1313. Robinson, B.E. et al., 2018: Incorporating land tenure security into conservation. Conserv. Lett., 11, e12383, doi:10.1111/conl.12383.
  1314. Deininger, K., and S. Jin, 2006: Tenure security and land-related investment: Evidence from Ethiopia. Eur. Econ. Rev., 50, 1245–1277, doi:10.1016/J.EUROECOREV.2005.02.001.
  1315. Etongo, D., I. Djenontin, M. Kanninen, and K. Fobissie, 2015: Smallholders’ tree planting activity in the Ziro Province, southern Burkina Faso: Impacts on livelihood and policy implications. Forests, 6, 2655–2677, doi:10.3390/f6082655.
  1316. Lawry, S., C. Samii, R. Hall, A. Leopold, D. Hornby, and F. Mtero, 2017: The impact of land property rights interventions on investment and agricultural productivity in developing countries: A systematic review. J. Dev. Eff., 9, 61–81, doi:10.1080/19439342.2016.1160947.
  1317. Abdulai, A., V. Owusu, and R. Goetz, 2011: Land tenure differences and investment in land improvement measures: Theoretical and empirical analyses. J. Dev. Econ., 96, 66–78, doi:10.1016/J.JDEVECO.2010.08.002.
  1318. Baumgartner, P., 2017: The Impacts of Large-Scale Land-Acquisition in East Africa on Poverty Reduction and the Rural Economy: Studies in Ethiopia and Uganda. PhD Thesis, University of Bonn, Bonn, Germany, 236 pp.
  1319. Dell’Angelo, J., P. D’Odorico, M.C. Rulli, and P. Marchand, 2017: The tragedy of the grabbed commons: Coercion and dispossession in the global land rush. World Dev., 92, 1–12, doi:10.1016/J.WORLDDEV.2016.11.005.
  1320. Basupi, L., C. Quinn, and A.J. Dougill, 2017: Pastoralism and land tenure transformation in Sub-Saharan Africa: Conflicting policies and priorities in Ngamiland, Botswana. Land, 6, 89, doi:10.3390/land6040089.
  1321. Kihiu, E., 2016a: Pastoral Practices, Economics, and Institutions of Sustainable Rangeland Management in Kenya. PhD Thesis, University of Bonn, Bonn, Germany, 167 pp.
  1322. Kihiu, E.N., 2016b: Basic capability effect: Collective management of pastoral resources in south-western Kenya. Ecol. Econ., 123, 23–34, doi:10.1016/J.ECOLECON.2016.01.003.
  1323. Behnke, R., 1994: Natural resource management in pastoral Africa. Dev. Policy Rev., 12, 5–28, doi:10.1111/j.1467-7679.1994.tb00053.x.
  1324. Holden, S.T., and H. Ghebru, 2016: Land tenure reforms, tenure security and food security in poor agrarian economies: Causal linkages and research gaps. Glob. Food Sec., 10, 21–28, doi:10.1016/J.GFS.2016.07.002.
  1325. Liao, C., P.E. Clark, S.D. DeGloria, and C.B. Barrett, 2017: Complexity in the spatial utilization of rangelands: Pastoral mobility in the Horn of Africa. Appl. Geogr., 86, 208–219, doi:10.1016/J.APGEOG.2017.07.003.
  1326. Turner, M.D., J.G. McPeak, K. Gillin, E. Kitchell, and N. Kimambo, 2016: Reconciling flexibility and tenure security for pastoral resources: The geography of transhumance networks in Eastern Senegal. Hum. Ecol., 44, 199–215, doi:10.1007/s10745-016-9812-2.
  1327. Wario, H.T., H.G. Roba, and B. Kaufmann, 2016: Responding to mobility constraints: Recent shifts in resource use practices and herding strategies in the Borana pastoral system, southern Ethiopia. J. Arid Environ., 127, 222–234, doi:10.1016/J.JARIDENV.2015.12.005.
  1328. Lambin, E.F. et al., 2014: Effectiveness and synergies of policy instruments for land use governance in tropical regions. Glob. Environ. Chang., 28, 129–140, doi:10.1016/J.GLOENVCHA.2014.06.007.
  1329. Reed, M.S., 2015: Reorienting land degradation towards sustainable land management: Linking sustainable livelihoods with ecosystem services in rangeland systems. J. Environ. Manage., 151, 472–485, doi:10.1016/J.JENVMAN.2014.11.010.
  1330. Schiappacasse, I., L. Nahuelhual, F. Vásquez, and C. Echeverría, 2012: Assessing the benefits and costs of Dryland forest restoration in central Chile. J. Environ. Manage., 97, 38–45. https://doi.org/10.1016/j.jenvman.2011.11.007.
  1331. Costanza, R., R. de Groot, and P. Sutton, 2014: Changes in the global value of ecosystem services. Glob. Environ. Chang., 26, 152–158, doi:10.1016/j.gloenvcha.2014.04.002.
  1332. Nkonya, E. et al., 2016a: Global cost of land degradation. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 117–165.
  1333. Börner, J., K. Baylis, E. Corbera, D. Ezzine-de-Blas, J. Honey-Rosés, U.M. Persson, and S. Wunder, 2017: The effectiveness of payments for environmental services. World Dev., 96, 359–374, doi:10.1016/J.WORLDDEV.2017.03.020.
  1334. Karsenty, A., and S. Ongolo, 2012: Can ‘fragile states’ decide to reduce their deforestation? The inappropriate use of the theory of incentives with respect to the REDD mechanism. For. Policy Econ., 18, 38–45, doi:10.1016/J.FORPOL.2011.05.006.
  1335. Gibson, C.C., J.T. Williams, and E. Ostrom, 2005: Local enforcement and better forests. World Dev., 33, 273–284, doi:10.1016/J.WORLDDEV.2004.07.013.
  1336. Smucker, T.A. et al., 2015: Differentiated livelihoods, local institutions, and the adaptation imperative: Assessing climate change adaptation policy in Tanzania. Geoforum, 59, 39–50, doi:10.1016/j.geoforum.2014.11.018.
  1337. Fernandez-Gimenez, M.E., B. Batkhishig, and B. Batbuyan, 2015: Lessons from the dzud: Community-based rangeland management increases the adaptive capacity of Mongolian herders to winter disasters. World Dev., 68, 48–65. doi:10.1016/j.worlddev.2014.11.015.
  1338. Abdu, N. and L. Robinson, 2017: Community-Based Rangeland Management in Dirre Rangeland Unit: Taking Successes in Land Restoration to Scale Project. ILRI Project Report. Nairobi, Kenya, 43 pp.
  1339. Dressler, W. et al., 2010: From hope to crisis and back again? A critical history of the global CBNRM narrative. Environ. Conserv., 37, 5–15, doi:10.1017/S0376892910000044.
  1340. Ostrom, E., and H. Nagendra, 2006: Insights on linking forests, trees, and people from the air, on the ground, and in the laboratory. Proc. Natl. Acad. Sci. U.S.A., 103, 19224–19231, doi:10.1073/pnas.0607962103.
  1341. Andela, N., Y.Y. Liu, A.I.J.M. van Dijk, R.A.M. de Jeu, and T.R. McVicar, 2013: Global changes in Dryland vegetation dynamics (1988-2008) assessed by satellite remote sensing: Comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences, 10, 6657–6676, doi:10.5194/bg-10-6657-2013.
  1342. Andersson, K.P., and E. Ostrom, 2008: Analyzing decentralized resource regimes from a polycentric perspective. Policy Sci., 41, 71–93, doi:10.1007/s11077-007-9055-6.
  1343. Cullman, G., 2015: Community forest management as virtualism in north-eastern Madagascar. Hum. Ecol., 43, 29–41, doi:10.1007/s10745-015-9725-5.
  1344. Dressler, W. et al., 2010: From hope to crisis and back again? A critical history of the global CBNRM narrative. Environ. Conserv., 37, 5–15, doi:10.1017/S0376892910000044.
  1345. Kadirbeyoglu, Z., and G. Özertan, 2015: Power in the governance of common-pool resources: A comparative analysis of irrigation management decentralization in Turkey. Environ. Policy Gov., 25, 157–171, doi:10.1002/eet.1673.
  1346. Umutoni, C., A. Ayantunde, M. Turner, and G.J. Sawadogo, 2016: Community participation in decentralized management of natural resources in the southern region of Mali. Environ. Nat. Resour. Res., 6, 1, doi:10.5539/enrr.v6n2p1.
  1347. Andela, N., Y.Y. Liu, A.I.J.M. van Dijk, R.A.M. de Jeu, and T.R. McVicar, 2013: Global changes in Dryland vegetation dynamics (1988-2008) assessed by satellite remote sensing: Comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences, 10, 6657–6676, doi:10.5194/bg-10-6657-2013.
  1348. Andersson, K.P., and E. Ostrom, 2008: Analyzing decentralized resource regimes from a polycentric perspective. Policy Sci., 41, 71–93, doi:10.1007/s11077-007-9055-6.
  1349. Carletto, C., P. Corral, and A. Guelfi, 2017: Agricultural commercialization and nutrition revisited: Empirical evidence from three African countries. Food Policy, 67, 106–118, doi:10.1016/J.FOODPOL.2016.09.020.
  1350. Carlisle, K., and R.L. Gruby, 2017: Polycentric systems of governance: A theoretical model for the commons. Policy Stud. J., doi:10.1111/psj.12212.
  1351. McCord, P., J. Dell’Angelo, E. Baldwin, and T. Evans, 2017: Polycentric transformation in Kenyan water governance: A dynamic analysis of institutional and social-ecological change. Policy Stud. J., 45, 633–658, doi:10.1111/psj.12168.
  1352. Çalişkan, S., and M. Boydak, 2017: Afforestation of arid and semi-arid ecosystems in Turkey. Turkish J. Agric. For., 41, 317–330, doi:10.3906/tar-1702-39.
  1353. Yurtoglu, M.A., 2015: The afforestation and erosion control mobilization action plan in Turkey. In: Living Land [Griffiths, J. (ed.)]. UNCCD and Tudor Rose, Leicester, United Kingdom, pp. 37–39.
  1354. Çalişkan, S., and M. Boydak, 2017: Afforestation of arid and semi-arid ecosystems in Turkey. Turkish J. Agric. For., 41, 317–330, doi:10.3906/tar-1702-39.
  1355. Turner, B.L., E.F. Lambin, and A. Reenberg, 2007: The emergence of land change science for global environmental change and sustainability. Proc. Natl. Acad. Sci., 104, 20666–20671, doi:10.1073/pnas.0704119104.
  1356. Grainger, A., 2009: The role of science in implementing international environmental agreements: The case of desertification. L. Degrad. Dev., 20, 410–430, doi:10.1002/ldr.898.
  1357. Harris, D., and A. Orr, 2014: Is rainfed agriculture really a pathway from poverty? Agric. Syst., 123, 84–96, doi:10.1016/J.AGSY.2013.09.005.
  1358. Akhtar-Schuster, M., R.J. Thomas, L.C. Stringer, P. Chasek, and M. Seely, 2011: Improving the enabling environment to combat land degradation: Institutional, financial, legal and science-policy challenges and solutions. L. Degrad. Dev., 22, 299–312, doi:10.1002/ldr.1058.
  1359. Verstraete, M.M. et al., 2011: Towards a global drylands observing system: Observational requirements and institutional solutions. L. Degrad. Dev., 22, 198–213, doi:10.1002/ldr.1046.
  1360. Chasek, P., W. Essahli, M. Akhtar-Schuster, L.C. Stringer, and R. Thomas, 2011: Integrated land degradation monitoring and assessment: Horizontal knowledge management at the national and international levels. L. Degrad. Dev., 22, 272–284, doi:10.1002/ldr.1096.
  1361. Reed, M.S., E. Nkonya, M. Winslow, M. Mortimore, and A. Mirzabaev, 2011: Monitoring and assessing the influence of social, economic and policy factors on sustainable land management in drylands. L. Degrad. Dev., 22, 240–247, doi:10.1002/ldr.1048.
  1362. Reed, M.S. et al., 2011: Cross-scale monitoring and assessment of land degradation and sustainable land management: A methodological framework for knowledge management. L. Degrad. Dev., 22, 261–271, doi:10.1002/ldr.1087.
  1363. Amugune, I., P.O. Cerutti, H. Baral, S. Leonard, and C. Martius, 2017: Small Flame but No Fire: Wood Fuel in the (Intended) Nationally Determined Contributions of Countries in Sub-Saharan Africa. Center for International Forestry Research (CIFOR), Bogor, Indonesia, 35 pp.
  1364. Iiyama, M., P. Dobie, M. Njenga, G. Ndegwa, and R. Jamnadass, 2014: The potential of agroforestry in the provision of sustainable woodfuel in Sub-Saharan Africa. Curr. Opin. Environ. Sustain., 6, 138–147, doi:10.1016/J.COSUST.2013.12.003.
  1365. Mekuria, W., M. Yami, M. Haile, K. Gebrehiwot, and E. Birhane, 2018: Impact of exclosures on wood biomass production and fuelwood supply in northern Ethiopia. J. For. Res., 30, 629–637, doi:10.1007/s11676-018-0643-4.
  1366. Neufeldt, H., K. Langford, J. Fuller, M. Iiyama, and P. Dobie, 2015: From Transition Fuel to Viable Energy Source: Improving Sustainability in the Sub-Saharan Charcoal Sector. ICRAF Working Paper No. 196, World Agroforestry Centre, Nairobi, Kenya, 20 pp.
  1367. Zulu, L.C., 2010: The forbidden fuel: Charcoal, urban woodfuel demand and supply dynamics, community forest management and woodfuel policy in Malawi. Energy Policy, 38, 3717–3730, doi:10.1016/J.ENPOL.2010.02.050.
  1368. Simon, G.L., and C. Peterson, 2018: Disingenuous forests: A historical political ecology of fuelwood collection in South India. J. Hist. Geogr., doi:10.1016/J.JHG.2018.09.003.
  1369. Swemmer, A.M., M. Mashele, and P.D. Ndhlovu, 2018: Evidence for ecological sustainability of fuelwood harvesting at a rural village in South Africa. Reg. Environ. Chang., 1–11, doi:10.1007/s10113-018-1402-y.
  1370. Twine, W.C., and R.M. Holdo, 2016: Fuelwood sustainability revisited: Integrating size structure and resprouting into a spatially realistic fuelshed model. J. Appl. Ecol., 53, 1766–1776, doi:10.1111/1365-2664.12713.
  1371. de la Sota, C., J. Lumbreras, N. Pérez, M. Ealo, M. Kane, I. Youm, and M. Viana, 2018: Indoor air pollution from biomass cookstoves in rural Senegal. Energy Sustain. Dev., 43, 224–234, doi:10.1016/J.ESD.2018.02.002.
  1372. Burke, P.J., and G. Dundas, 2015: Female labour force participation and household dependence on biomass energy: Evidence from National Longitudinal Data. World Dev., 67, 424–437, doi:10.1016/J.WORLDDEV.2014.10.034.
  1373. Alba, R., A. Bolding, and R. Ducrot, 2016: The politics of water payments and stakeholder participation in the Limpopo River Basin, Mozambique Raphaëlle Ducrot. Water Altern., 9, 569–587.
  1374. Albanito, F. et al., 2016: Carbon implications of converting cropland to bioenergy crops or forest for climate mitigation: A global assessment. GCB Bioenergy, 8, 81–95, doi:10.1111/gcbb.12242.
  1375. Nair, P.K.R., S.K. Saha, V.D. Nair, and S.G. Haile, 2011: Potential for greenhouse gas emissions from soil carbon stock following biofuel cultivation on degraded lands. L. Degrad. Dev., 22, 395–409, doi:10.1002/ldr.1016.
  1376. Negash, D., A. Abegaz, J.U. Smith, H. Araya, and B. Gelana, 2017: Household energy and recycling of nutrients and carbon to the soil in integrated crop-livestock farming systems: A case study in Kumbursa village, Central Highlands of Ethiopia. GCB Bioenergy, 9, 1588–1601, doi:10.1111/gcbb.12459.
  1377. Bang, G., 2010: Energy security and climate change concerns: Triggers for energy policy change in the United States? Energy Policy, 38, 1645–1653, doi:10.1016/J.ENPOL.2009.01.045.
  1378. El-Fadel, M., R. Chedid, M. Zeinati, and W. Hmaidan, 2003: Mitigating energy-related GHG emissions through renewable energy. Renew. Energy, 28, 1257–1276, doi:10.1016/S0960-1481(02)00229-X.
  1379. Cancino-Solórzano, Y., J.P. Paredes-Sánchez, A.J. Gutiérrez-Trashorras, and 
J. Xiberta-Bernat, 2016: The development of renewable energy resources in the State of Veracruz, Mexico. Util. Policy, 39, 1–4, doi:10.1016/J.JUP.2016.01.001.
  1380. Hong, L., N. Zhou, D. Fridley, and C. Raczkowski, 2013: Assessment of China’s renewable energy contribution during the 12th Five Year Plan. Energy Policy, 62, 1533–1543, doi:10.1016/J.ENPOL.2013.07.110.
  1381. Farnoosh, A., F. Lantz, and J. Percebois, 2014: Electricity generation analyses in an oil-exporting country: Transition to non-fossil fuel based power units in Saudi Arabia. Energy, 69, 299–308, doi:10.1016/J.ENERGY.2014.03.017.
  1382. Stambouli, A.B., Z. Khiat, S. Flazi, and Y. Kitamura, 2012: A review on the renewable energy development in Algeria: Current perspective, energy scenario and sustainability issues. Renew. Sustain. Energy Rev., 16, 4445–4460, doi:10.1016/j.rser.2012.04.031.
  1383. Vidadili, N., E. Suleymanov, C. Bulut, and C. Mahmudlu, 2017: Transition to renewable energy and sustainable energy development in Azerbaijan. Renew. Sustain. Energy Rev., 80, 1153–1161, doi:10.1016/J.RSER.2017.
05.168.
  1384. Baker, L., P. Newell, and J. Phillips, 2014: The political economy of energy transitions: The case of South Africa. New Polit. Econ., 19, 791–818, doi:10.1080/13563467.2013.849674.
  1385. Elum, Z. and A.S. Momodu, 2017: Climate change mitigation and renewable energy for sustainable development in Nigeria: A discourse approach. Renew. Sustain. Energy Rev., 76, 72–80.
  1386. Karatayev, M., S. Hall, Y. Kalyuzhnova, and M.L. Clarke, 2016: Renewable energy technology uptake in Kazakhstan: Policy drivers and barriers in a transitional economy. Renew. Sustain. Energy Rev., 66, 120–136, doi:10.1016/J.RSER.2016.07.057.
  1387. Aliyu, A., B. Modu, and C.W. Tan, 2017: A review of renewable energy development in Africa: A focus in South Africa, Egypt and Nigeria. Renew. Sustain. Energy Rev., 81, 2502–2518, doi:10.1016/j.rser.2017.06.055.
  1388. Vogt, J.V., U. Safriel, G. Von Maltitz, Y. Sokona, R. Zougmore, G. Bastin, and J. Hill, 2011: Monitoring and assessment of land degradation and desertification: Towards new conceptual and integrated approaches. L. Degrad. Dev., 22, 150–165, doi:10.1002/ldr.1075.
  1389. Pozzi, W. et al., 2013: Toward global drought early warning capability: Expanding international cooperation for the development of a framework for monitoring and forecasting. Bull. Am. Meteorol. Soc., 94, 776–785, doi:10.1175/BAMS-D-11-00176.1.
  1390. Bergametti, G. et al., 2018: Size-resolved dry deposition velocities of dust particles: In situ measurements and parameterizations testing. J. Geophys. Res. Atmos., 123, 11, 11–80, 99, doi:10.1029/2018JD028964.
  1391. Marticorena, B. et al., 2010: Temporal variability of mineral dust concentrations over West Africa: Analyses of a pluriannual monitoring from the AMMA Sahelian Dust Transect. Atmos. Chem. Phys., 10, 8899–8915, doi:10.5194/acp-10-8899-2010.
  1392. Beegum, S.N. et al., 2016: Simulating aerosols over Arabian Peninsula with CHIMERE: Sensitivity to soil, surface parameters and anthropogenic emission inventories. Atmos. Environ., 128, 185–197, doi:10.1016/J.ATMOSENV.2016.01.010.
  1393. Cornell, S. et al., 2013: Opening up knowledge systems for better responses to global environmental change. Environ. Sci. Policy, 28, 60–70, doi:10.1016/J.ENVSCI.2012.11.008.
  1394. Cornet, A., 2012: Surveillance environnementale et développement Des observations écologiques à la surveillance environnementale: un besoin pour comprendre et agir. Options Méditerranéennes, B, 1–14.
  1395. Haase, P. et al., 2018: The next generation of site-based long-term ecological monitoring: Linking essential biodiversity variables and ecosystem integrity. Sci. Total Environ., 613–614, 1376–1384, doi:10.1016/J.SCITOTENV.2017.08.111.
  1396. Sergeant, C.J., B.J. Moynahan, and W.F. Johnson, 2012: Practical advice for implementing long-term ecosystem monitoring. J. Appl. Ecol., 49, 969–973, doi:10.1111/j.1365-2664.2012.02149.x.
  1397. Cherlet, M. et al. (eds.), 2018: World Atlas of Desertification. Publication Office of the European Union, Luxembourg, 248 pp.
  1398. Geerken, R.A., 2009: An algorithm to classify and monitor seasonal variations in vegetation phenologies and their inter-annual change. ISPRS J. Photogramm. Remote Sens., 64, 422–431, doi:10.1016/J.ISPRSJPRS.2009.03.001.
  1399. Geerts, S., and D. Raes, 2009: Deficit irrigation as an on-farm strategy to maximize crop water productivity in dry areas. Agric. Water Manag., 96, 1275–1284, doi:10.1016/J.AGWAT.2009.04.009.
  1400. Olayide, O.E., I.K. Tetteh, and L. Popoola, 2016: Differential impacts of rainfall and irrigation on agricultural production in Nigeria: Any lessons for climate-smart agriculture? Agric. Water Manag., 178, 30–36, doi:10.1016/J.AGWAT.2016.08.034.
  1401. Oweis, T., and A. Hachum, 2006: Water harvesting and supplemental irrigation for improved water productivity of dry farming systems in West Asia and North Africa. Agric. Water Manag., 80, 57–73, doi:10.1016/J.AGWAT.2005.07.004.
  1402. Nkonya, E, T. Johnson, H.Y. Kwon, and E. Kato, 2016b: Economics of land degradation in Sub-Saharan Africa. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev, and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 215–259.
  1403. Altchenko, Y., and K.G. Villholth, 2015: Mapping irrigation potential from renewable groundwater in Africa – A quantitative hydrological approach. Hydrol. Earth Syst. Sci., 19, 1055–1067, doi:10.5194/hess-19-1055-2015.
  1404. Cherlet, M. et al. (eds.), 2018: World Atlas of Desertification. Publication Office of the European Union, Luxembourg, 248 pp.
  1405. Bjornlund, H., A. van Rooyen, and R. Stirzaker, 2017: Profitability and productivity barriers and opportunities in small-scale irrigation schemes. Int. J. Water Resour. Dev., 33, 690–704, doi:10.1080/07900627.2016.1263552.
  1406. Woodhouse, P., G.J. Veldwisch, J.-P. Venot, D. Brockington, H. Komakech, and Â. Manjichi, 2017: African farmer-led irrigation development: Re-framing agricultural policy and investment? J. Peasant Stud., 44, 213–233, doi:10.1080/03066150.2016.1219719.
  1407. Bazza, M., M. Kay, and C. Knutson, 2018: Drought Characteristics and Management in North Africa and the Near East. FAO Water Reports 45, Food and Agriculture Organization of the United Nations, Rome, Italy, 1020–1203 pp.
  1408. Kingdom of Saudi Arabia, 2016: National Transformation Program 2020, Vision 2030: Delivery Plan 2018–2020. Government of Saudi Arabia, Riyad, Saudi Arabia, pp. 1–57.
  1409. Behnke, R., and C. Kerven, 2013: Counting the costs: Replacing pastoralism with irrigated agriculture in the Awash Valley, north-eastern Ethiopia. In: Pastoralism and Development in Africa: Dynamic Change at the Margins [Catley, A., J. Lind, and I. Scoones (eds.)]. Routledge, London, UK, pp. 328.
  1410. Chandel, S.S., M. Nagaraju Naik, and R. Chandel, 2015: Review of solar photovoltaic water pumping system technology for irrigation and community drinking water supplies. Renew. Sustain. Energy Rev., 49, 1084–1099, doi:10.1016/J.RSER.2015.04.083.
  1411. Burney, J., L. Woltering, M. Burke, R. Naylor, and D. Pasternak, 2010: Solar-powered drip irrigation enhances food security in the Sudano-Sahel. Proc. Natl. Acad. Sci. U.S.A., 107, 1848–1853, doi:10.1073/pnas.0909678107.
  1412. Closas, A., and E. Rap, 2017: Solar-based groundwater pumping for irrigation: Sustainability, policies, and limitations. Energy Policy, 104, 33–37, doi:10.1016/J.ENPOL.2017.01.035.
  1413. Barrett, C.B., L. Christiaensen, M. Sheahan, and A. Shimeles, 2017: On the structural transformation of rural Africa. J. Afr. Econ., 26, i11–i35, doi:10.1093/jae/ejx009.
  1414. Awotide, B.A., A.A. Karimov, and A. Diagne, 2016: Agricultural technology adoption, commercialization and smallholder rice farmers’ welfare in rural Nigeria. Agric. Food Econ., 4, 3, doi:10.1186/s40100-016-0047-8.
  1415. Carletto, C., P. Corral, and A. Guelfi, 2017: Agricultural commercialization and nutrition revisited: Empirical evidence from three African countries. Food Policy, 67, 106–118, doi:10.1016/J.FOODPOL.2016.09.020.
  1416. Carlisle, K., and R.L. Gruby, 2017: Polycentric systems of governance: A theoretical model for the commons. Policy Stud. J., doi:10.1111/psj.12212.
  1417. Reardon, T. et al., 2009: Agrifood Industry Transformation and Small Farmers in Developing Countries. World Dev., 37, 1717–1727, doi:10.1016/J.WORLDDEV.2008.08.023.
  1418. Cour, J.M., 2001: The Sahel in West Africa: Countries in transition to a full market economy. Glob. Environ. Chang., 11, 31–47, doi:10.1016/S0959-3780(00)00043-1.
  1419. Reardon, T. et al., 2009: Agrifood Industry Transformation and Small Farmers in Developing Countries. World Dev., 37, 1717–1727, doi:10.1016/J.WORLDDEV.2008.08.023.
  1420. Haggblade, S., P. Hazell, and T. Reardon, 2010: The rural non-farm economy: Prospects for growth and poverty reduction. World Dev., 38, 1429–1441, doi:10.1016/J.WORLDDEV.2009.06.008.
  1421. Shiferaw, B., and A. Djido, 2016: Patterns of Labor Productivity and Income Diversification in the Rural Farm and Non-farm Sectors in Sub-Saharan Africa: Partnership for Economic Policy. Policy Brief No. 143. Nairobi, Kenya, 3 pp.
  1422. Shiferaw, B., and A. Djido, 2016: Patterns of Labor Productivity and Income Diversification in the Rural Farm and Non-farm Sectors in Sub-Saharan Africa: Partnership for Economic Policy. Policy Brief No. 143. Nairobi, Kenya, 3 pp.
  1423. Headey, D., A.S. Taffesse, and L. You, 2014: Diversification and development in pastoralist Ethiopia. World Dev., 56, 200–213, doi:10.1016/J.WORLDDEV.2013.10.015.
  1424. Liao, C., C. Barrett, and K.-A. Kassam, 2015: Does diversification improve livelihoods? Pastoral households in Xinjiang, China. Dev. Change, 46, 1302–1330, doi:10.1111/dech.12201.
  1425. Jedwab, R., and D. Vollrath, 2015: Urbanization without growth in historical perspective. Explor. Econ. Hist., 58, 1–21, doi:10.1016/J.EEH.2015.09.002.
  1426. Revi, A., and C. Rosenzweig, 2013: The Urban Opportunity to Enable Transformative and Sustainable Development. Background paper for the High-Level Panel of Eminent Persons on the Post-2015 Development Agenda. Prepared by the Co-Chairs of the Sustainable Development Solutions Network Thematic Group on Sustainable Cities. Paris, France and New York, USA, 47 pp.
  1427. Júnior, W.S.F., F.R. Santoro, I. Vandebroek, and U.P. Albuquerque, 2016: Urbanization, modernization, and nature knowledge. In: Introduction to Ethnobiology [Albuquerque, U. and R. Nobrega Alves (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 251–256.
  1428. Reyer, C.P.O. et al., 2013: A plant’s perspective of extremes: Terrestrial plant responses to changing climatic variability. Glob. Chang. Biol., 19, 75–89, doi:10.1111/gcb.12023.
  1429. Reyes-García, V., M. Guèze, A.C. Luz, J. Paneque-Gálvez, M.J. Macía, 
M. Orta-Martínez, J. Pino, and X. Rubio-Campillo, 2013: Evidence of traditional knowledge loss among a contemporary indigenous society. Evol. Hum. Behav., 34, 249–257, doi:10.1016/J.EVOLHUMBEHAV.2013.03.002.
  1430. van Andel, T., and L.G. Carvalheiro, 2013: Why urban citizens in developing countries use traditional medicines: The case of Suriname. Evid. Based. Complement. Alternat. Med., 2013, 687197, doi:10.1155/2013/687197.
  1431. Angel, S., J. Parent, D.L. Civco, and A.M. Blei, 2011: Making Room for a Planet of Cities. Lincoln Institute of Land Policy, Cambridge, USA, 76 pp.
  1432. Cour, J.M., 2001: The Sahel in West Africa: Countries in transition to a full market economy. Glob. Environ. Chang., 11, 31–47, doi:10.1016/S0959-3780(00)00043-1.
  1433. Dahiya, B., 2012: Cities in Asia, 2012: Demographics, economics, poverty, environment and governance. Cities, 29, S44–S61, doi:10.1016/J.CITIES.2012.06.013.
  1434. Bulkeley, H., 2013: Cities and Climate Change. Routledge, London, UK, 268 pp.
  1435. Garschagen, M., and P. Romero-Lankao, 2015: Exploring the relationships between urbanization trends and climate change vulnerability. Clim. Change, 133, 37–52, doi:10.1007/s10584-013-0812-6.
  1436. Wilhite, D., and R.S. Pulwarty, 2017: Drought and Water Crises, Integrating Science, Management, and Policy, Second Edition. CRC Press, Boca Raton, Florida, USA, 542 pp.
  1437. Ford, T.W., and C.F. Labosier, 2017: Meteorological conditions associated with the onset of flash drought in the Eastern United States. Agric. For. Meteorol., 247, 414–423, doi:10.1016/J.AGRFORMET.2017.08.031.
  1438. Wilhite, D., and R.S. Pulwarty, 2017: Drought and Water Crises, Integrating Science, Management, and Policy, Second Edition. CRC Press, Boca Raton, Florida, USA, 542 pp.
  1439. Cook, B.I., Smerdon, J.E.,, R. Seager, and E.R. Cook, 2014b: Pan-continental droughts in North America over the last millennium. J. Clim., 27, 383–397, doi:10.1175/JCLI-D-13-00100.1.
  1440. Seneviratne, S.I., and P. Ciais, 2017: Environmental science: Trends in ecosystem recovery from drought. Nature, 548, 164–165, doi:10.1038/548164a.
  1441. Spinoni, J. et al., 2019: A new global database of meteorological drought events from 1951 to 2016. J. Hydrol. Reg. Stud., 22, 100593, doi:10.1016/J.EJRH.2019.100593.
  1442. Türkeş. M., 1999: Vulnerability of Turkey to desertification with respect to precipitation and aridity conditions. Turkish J. Eng. Environ. Sci., 23, 363–380.
  1443. Wilhite, D.A., M.V.K. Sivakumar, and R. Pulwarty, 2014: Managing drought risk in a changing climate: The role of national drought policy. Weather Clim. Extrem., 3, 4–13, doi:10.1016/J.WACE.2014.01.002.
  1444. Cook, B.I., R.L. Miller, and R. Seager, 2009: Amplification of the North American ‘Dust Bowl’ drought through human-induced land degradation. Proc. Natl. Acad. Sci. U.S.A., 106, 4997–5001, doi:10.1073/pnas.0810200106.
  1445. Hornbeck, R., 2012: The enduring impact of the American Dust Bowl: Short- and long-run adjustments to environmental catastrophe. Am. Econ. Rev., 102, 1477–1507, doi:10.1257/aer.102.4.1477.
  1446. Gerber, N., and A. Mirzabaev, 2017: Benefits of Action and Costs of Inaction: Drought Mitigation and Preparedness – A Literature Review. Integrated Drought Management Programme Working Paper No. 1., World Meteorological Organization, Geneva, Switzerland and Global Water Partnership, Stockholm, Sweden, 23 pp.
  1447. CRED, 2015: Human Cost of Natural Disasters 2015: A global perspective. CRED and UNISDR. Brussels, Belgium. 59 pp.
  1448. Winsemius, H.C., B. Jongman, T.I.E. Veldkamp, S. Hallegatte, M. Bangalore, and P.J. Ward, 2018: Disaster risk, climate change, and poverty: Assessing the global exposure of poor people to floods and droughts. Environ. Dev. Econ., 23, 328–348, doi:10.1017/S1355770X17000444.
  1449. Mosley, L., 2014: Drought impacts on the water quality of freshwater systems; Review and integration. Earth-Science Rev., 140, 203–214, doi:10.1016/j.earscirev.2014.11.010.
  1450. Gray, C., and V. Mueller, 2012a: Drought and population mobility in rural Ethiopia. World Dev., 40, 134–145, doi:10.1016/J.WORLDDEV.2011.05.023.
  1451. Gray, C.L., and V. Mueller, 2012b: Natural disasters and population mobility in Bangladesh. Proc. Natl. Acad. Sci. U.S.A., 109, 6000–6005, doi:10.1073/pnas.1115944109.
  1452. Johnstone, S., and J. Mazo, 2011: Global warming and the Arab Spring. Survival (Lond)., 53, 11–17, doi:10.1080/00396338.2011.571006.
  1453. Linke, A.M., J. O’Loughlin, J.T. McCabe, J. Tir, and F.D.W. Witmer, 2015: Rainfall variability and violence in rural Kenya: Investigating the effects of drought and the role of local institutions with survey data. Glob. Environ. Chang., 34, 35–47, doi:10.1016/J.GLOENVCHA.2015.04.007.
  1454. Lohmann, S., and T. Lechtenfeld, 2015: The effect of drought on health outcomes and health expenditures in rural Vietnam. World Dev., 72, 432–448, doi:10.1016/J.WORLDDEV.2015.03.003.
  1455. Maystadt, J.-F., and O. Ecker, 2014: Extreme weather and civil war: Does drought fuel conflict in Somalia through livestock price shocks? Am. J. Agric. Econ., 96, 1157–1182, doi:10.1093/ajae/aau010.
  1456. Yusa, A. et al., 2015: Climate change, drought and human health in Canada. Int. J. Environ. Res. Public Health, 12, 8359–8412, doi:10.3390/ijerph120708359.
  1457. Botterill, L.C., and M.J. Hayes, 2012: Drought triggers and declarations: Science and policy considerations for drought risk management. Nat. Hazards, 64, 139–151, doi:10.1007/s11069-012-0231-4.
  1458. Gerber, N., and A. Mirzabaev, 2017: Benefits of Action and Costs of Inaction: Drought Mitigation and Preparedness – A Literature Review. Integrated Drought Management Programme Working Paper No. 1., World Meteorological Organization, Geneva, Switzerland and Global Water Partnership, Stockholm, Sweden, 23 pp.
  1459. Clarke, D.J., and R.V. Hill, 2013: Cost-Benefit Analysis of the African Risk Capacity Facility. No. 01292, Washington, DC, USA, 64 pp.
  1460. Visser, W.P., 2018: A perfect storm: The ramifications of Cape Town’s drought crisis. J. Transdiscipl. Res. South. Africa, 14, 1–10, doi:10.4102/td.v14i1.567.
  1461. Sivakumar, M.V.K., Ndiang’ui, N. (ed.), 2007: Climate and Land Degradation (Environmental Science and Engineering). Springer-Verlag, Berlin, Germany, 623 pp.
  1462. Sivakumar, M.V.K., 2007: Interactions between climate and desertification. Agric. For. Meteorol., 142, 143–155, doi:10.1016/j.agrformet.2006.03.025.
  1463. Verbist, K., A. Amani, A. Mishra, and B. Jiménez, 2016: Strengthening drought risk management and policy: UNESCO International Hydrological Programme’s case studies from Africa and Latin America and the Caribbean. Water Policy, 18, 245–261, doi:10.2166/wp.2016.223.
  1464. Hallegatte, S., 2012: A Cost Effective Solution to Reduce Disaster Losses in Developing Countries: Hydro-Meteorological Services, Early Warning, and Evacuation. Policy Research Working Paper no. WPS 6058, The World Bank, Washington DC, USA, 22 pp.
  1465. Guerrero-Baena, M., J. Gómez-Limón, M.D. Guerrero-Baena, and 
J.A. Gómez-Limón, 2019: Insuring water supply in irrigated agriculture: 
A proposal for hydrological drought index-based insurance in Spain. Water, 11, 686, doi:10.3390/w11040686.
  1466. Kath, J., S. Mushtaq, R. Henry, A.A. Adeyinka, R. Stone, T. Marcussen, and 
L. Kouadio, 2019: Spatial variability in regional scale drought index insurance viability across Australia’s wheat growing regions. Clim. Risk Manag., 24, 13–29, doi:10.1016/J.CRM.2019.04.002.
  1467. Osgood, D. et al., 2018: Farmer perception, recollection, and remote sensing in weather index insurance: An Ethiopia case study. Remote Sens., 10, 1887, doi:10.3390/rs10121887.
  1468. Ruiz, J., M. Bielza, A. Garrido, and A. Iglesias, 2015: Dealing with drought in irrigated agriculture through insurance schemes: An application to an irrigation district in southern Spain. Spanish J. Agric. Res., 13, e0106, doi:10.5424/sjar/2015134-6941.
  1469. Tadesse, M.A., B.A. Shiferaw, and O. Erenstein, 2015: Weather index insurance for managing drought risk in smallholder agriculture: Lessons and policy implications for Sub-Saharan Africa. Agric. Food Econ., 3, 26, doi:10.1186/s40100-015-0044-3.
  1470. Vicente-Serrano, S.M. et al., 2012: Challenges for drought mitigation in Africa: The potential use of geospatial data and drought information systems. Appl. Geogr., 34, 471–486, doi:10.1016/J.APGEOG.2012.02.001.
  1471. Tsakiris, G., 2017: Facets of modern water resources management: Prolegomena. Water Resour. Manag., 31, 2899–2904, doi:10.1007/s11269-017-1742-2.
  1472. Yusa, A. et al., 2015: Climate change, drought and human health in Canada. Int. J. Environ. Res. Public Health, 12, 8359–8412, doi:10.3390/ijerph120708359.
  1473. Booysen, M.J., M. Visser, and R. Burger, 2019: Temporal case study of household behavioural response to Cape Town’s ‘Day Zero’ using smart meter data. Water Res., 149, 414–420, doi:10.1016/J.WATRES.2018.11.035.
  1474. Salami, H., N. Shahnooshi, and K.J. Thomson, 2009: The economic impacts of drought on the economy of Iran: An integration of linear programming and macroeconometric modelling approaches. Ecol. Econ., 68, 1032–1039, doi:10.1016/J.ECOLECON.2008.12.003.
  1475. Kath, J., S. Mushtaq, R. Henry, A.A. Adeyinka, R. Stone, T. Marcussen, and 
L. Kouadio, 2019: Spatial variability in regional scale drought index insurance viability across Australia’s wheat growing regions. Clim. Risk Manag., 24, 13–29, doi:10.1016/J.CRM.2019.04.002.
  1476. Hurlbert, M.A., 2018: Adaptive Governance of Disaster: Drought and Flood in Rural Areas. Springer International Publishing, Springer. Cham, Switzerland. ISBN 978-3-319-57801-9, 239 pp.
  1477. Hurlbert, M. and P. Mussetta, 2016: Creating resilient water governance for irrigated producers in Mendoza, Argentina. Environ. Sci. Policy, 58, 83–94, doi:10.1016/j.envsci.2016.01.004.
  1478. Hurlbert, M.A., 2018: Adaptive Governance of Disaster: Drought and Flood in Rural Areas. Springer International Publishing, Springer. Cham, Switzerland. ISBN 978-3-319-57801-9, 239 pp.
  1479. Alam, K., 2015: Farmers’ adaptation to water scarcity in drought-prone environments: A case study of Rajshahi District, Bangladesh. Agric. Water Manag., 148, 196–206, doi:10.1016/J.AGWAT.2014.10.011.
  1480. Kusunose, Y., and T.J. Lybbert, 2014: Coping with drought by adjusting land tenancy contracts: A model and evidence from rural Morocco. World Dev., 61, 114–126, doi:10.1016/J.WORLDDEV.2014.04.006.
  1481. Verner, D. et al., 2018: Climate Variability, Drought, and Drought Management in Morocco’s Agricultural Sector. World Bank Group, Washington, DC, USA. https://doi.org/10.1596/30603, 146 pp.
  1482. Wilhite, D.A., 2016: Drought-management policies and preparedness plans: Changing the paradigm from crisis to risk management. L. Restor., 443–462, doi:10.1016/B978-0-12-801231-4.00007-0.
  1483. Jerneck, A., and L. Olsson, 2008: Adaptation and the poor: Development, resilience and transition. Clim. Policy, 8, 170–182, doi:10.3763/cpol.2007.0434.
  1484. McLeman, R., 2013: Developments in modelling of climate change-related migration. Clim. Change, 117, 599–611, doi:10.1007/s10584-012-0578-2.
  1485. Wilhite, D.A., M.V.K. Sivakumar, and R. Pulwarty, 2014: Managing drought risk in a changing climate: The role of national drought policy. Weather Clim. Extrem., 3, 4–13, doi:10.1016/J.WACE.2014.01.002.
  1486. Barnett, J., L.S. Evans, C. Gross, A.S. Kiem, R.T. Kingsford, J.P. Palutikof, 
C.M. Pickering, and S.G. Smithers, 2015: From barriers to limits to climate change adaptation: path dependency and the speed of change. Ecol. Soc., 20, 5, doi:10.5751/ES-07698-200305.
  1487. Eekhout, J. and J. de Vente, 2019: Assessing the effectiveness of Sustainable Land Management for large-scale climate change adaptation. Sci. Total Environ., 654, 85–93, doi:10.1016/J.SCITOTENV.2018.10.350.
  1488. Dixon, J.A., D.E. James, P.B. Sherman, D.E. James, and P.B. Sherman, 2013: Economics of Dryland Management. Routledge, London, UK. 324 pp.
  1489. Waha, K., C. Müller, A. Bondeau, J.P. Dietrich, P. Kurukulasuriya, J. Heinke, and H. Lotze-Campen, 2013: Adaptation to climate change through the choice of cropping system and sowing date in Sub-Saharan Africa. Glob. Environ. Chang., 23, 130–143, doi:10.1016/J.GLOENVCHA.2012.11.001.
  1490. Kamali, B., K.C. Abbaspour, B. Wehrli, and H. Yang, 2018: Drought vulnerability assessment of maize in Sub-Saharan Africa: Insights from physical and social perspectives. Glob. Planet. Change, 162, 266–274, doi:10.1016/J.GLOPLACHA.2018.01.011.
  1491. Schneider, F.D., and S. Kéfi, 2016: Spatially heterogeneous pressure raises risk of catastrophic shifts. Theor. Ecol., 9, 207–217, doi:10.1007/s12080-015-0289-1.
  1492. Godfray, H.C.J., and T. Garnett, 2014: Food security and sustainable intensification. Philos. Trans. R. Soc. B Biol. Sci., 369, 20120273, doi:10.1098/rstb.2012.0273.
  1493. Stringer, L.C., et al., 2012: Challenges and opportunities for carbon management in Malawi and Zambia. Carbon Manag., 3, 159–173, doi:10.4155/cmt.12.14.
  1494. Guodaar, L. et al., 2019: How do climate change adaptation strategies result in unintended maladaptive outcomes? Perspectives of tomato farmers. Int. J. Veg. Sci., 1–17, doi:10.1080/19315260.2019.1573393.
  1495. Müller, B., L. Johnson, and D. Kreuer, 2017: Maladaptive outcomes of climate insurance in agriculture. Glob. Environ. Chang., 46, 23–33, doi:10.1016/J.GLOENVCHA.2017.06.010.
  1496. Al Hamndou, D., and M. Requier-Desjardins, 2008: Variabilité climatique, désertification et biodiversité en afrique: S’adapter, une approche intégrée. VertigO, 8(2): 1-24, doi:10.4000/vertigo.5356.
  1497. Antwi-Agyei, P., A.J. Dougill, L.C. Stringer, and S.N.A. Codjoe, 2018: Adaptation opportunities and maladaptive outcomes in climate vulnerability hotspots of Northern Ghana. Clim. Risk Manag., 19, 83–93, doi:10.1016/J.CRM.2017.11.003.
  1498. Antwi-Agyei, P., A.J. Dougill, L.C. Stringer, and S.N.A. Codjoe, 2018: Adaptation opportunities and maladaptive outcomes in climate vulnerability hotspots of Northern Ghana. Clim. Risk Manag., 19, 83–93, doi:10.1016/J.CRM.2017.11.003.
  1499. Apuri, I., K. Peprah, and G.T.W. Achana, 2018: Climate change adaptation through agroforestry: The case of Kassena Nankana West District, Ghana. Environ. Dev., 28, 32–41, doi:10.1016/J.ENVDEV.2018.09.002.
  1500. Achite, M., and S. Ouillon, 2007: Suspended sediment transport in a semi-arid watershed, Wadi Abd, Algeria (1973–1995). J. Hydrol., 343, 187–202, doi:10.1016/J.JHYDROL.2007.06.026.
  1501. Megnounif, A. and A.N. Ghenim, 2016: Rainfall irregularity and its impact on the sediment yield in Wadi Sebdou watershed, Algeria. Arab. J. Geosci., 9, 267, doi:10.1007/s12517-015-2280-y.
  1502. Vachtman, D., A. Sandler, N. Greenbaum, and B. Herut, 2013: Dynamics of suspended sediment delivery to the Eastern Mediterranean continental shelf. Hydrol. Process., 27, 1105–1116, doi:10.1002/hyp.9265.
  1503. Zhang, X.C., and M.A. Nearing, 2005: Impact of climate change on soil erosion, runoff, and wheat productivity in central Oklahoma. CATENA, 61, 185–195, doi:10.1016/J.CATENA.2005.03.009.
  1504. Anache, J.A.A., D.C. Flanagan, A. Srivastava, and E.C. Wendland, 2018: Land use and climate change impacts on runoff and soil erosion at the hillslope scale in the Brazilian Cerrado. Sci. Total Environ., 622–623, 140–151, doi:10.1016/J.SCITOTENV.2017.11.257.
  1505. Poesen, J., 2018: Soil erosion in the Anthropocene: Research needs. Earth Surf. Process. Landforms, 43, 64–84, doi:10.1002/esp.4250.
  1506. Ziadat, F.M., and A.Y. Taimeh, 2013: Effect of rainfall intensity, slope, land use and antecedent soil moisture on soil erosion in an arid environment. L. Degrad. Dev., 24, 582–590, doi:10.1002/ldr.2239.
  1507. García-Ruiza, J.M., S. Beguería, N. Lana-Renault, E. Nadal-Romero, and 
A. Cerdà, 2017: Ongoing and emerging questions in water erosion studies. L. Degrad. Dev., 28, 5–21, doi:10.1002/ldr.2641.
  1508. CIREN, 2010: Determinación de la erosión potencial y actual de los suelos de Chile. Centro de Información de Recursos Naturales, Santiago, Chile, 285 pp.
  1509. Bonilla, C.A., J.L. Reyes, and A. Magri, 2010: Water erosion prediction using the revised universal soil loss equation (RUSLE) in a GIS framework, Central Chile. Chil. J. Agric. Res., 70, 159–169, doi:10.4067/S0718-58392010000100017.
  1510. Reicosky, D., and C. Crovetto, 2014: No-till systems on the Chequen Farm in Chile: A success story in bringing practice and science together. Int. Soil Water Conserv. Res., 2, 66–77, doi:10.1016/S2095-6339(15)30014-9.
  1511. Martínez, G.I. et al., 2011: Influence of conservation tillage and soil water content on crop yield in Dryland compacted Alfisol of Central Chile. Chil. J. Agric. Res., 71, 615–622, doi:10.4067/S0718-58392011000400018.
  1512. Acevedo, E., and P. Silva, 2003: Agronomía de la cero labranza. Ser. Ciencias Agronómicas, 3, 118. Universidad de Chile, ISBN: 956-19-0363-6.
  1513. Türkeş, M., 2003: Spatial and temporal variations in precipitation and aridity index series of Turkey. In: Mediterranean Climate [Bolle, H.J. (ed.)]. Springer, Berlin, Germany, pp. 181–213.
  1514. Türkeş, M., and H. Tatlı, 2011: Use of the spectral clustering to determine coherent precipitation regions in Turkey for the period 1929–2007. Int. J. Climatol., 31, 2055–2067, doi:10.1002/joc.2212.
  1515. Groneman, A., 1968: The Soils of the Wind Erosion Control Camp Area Karapinar, Turkey. Publication Series No. 472, Wageningen Agricultural University, Wageningen, Netherlands, 161 pp.
  1516. Kantarcı, M.D., Özel, H.B., Ertek, M. and Kırdar, E., 2011: An assessment on the adaptation of six tree species to steppe habitat during Konya-Karapinar sand-dune afforestations. J. Bartın For. Fac., 13, 107–127.
  1517. Akay, A., and A.I. Yildirim, 2010: The Greening Desert Of Karapinar: An Example from Turkey. Soil and Water Resources Research Institute Laboratory Agricultural Engineer, Konya, Turkey, pp. 546–551.
  1518. Akça, E., K. Takashi, and T. Sato, 2016: Development and success, for whom and where: The central Anatolian case. IN: Chabay I, Frick M, Helgeson J. Land Restoration.Pages 33–541, Boston, Academic Press, doi:10.1016/B978-0-12-801231-4.00034-3.
  1519. Saparov, A., 2014: Soil resources of the Republic of Kazakhstan: Current status, problems and solutions. In: Novel Measurement and Assessment Tools for Monitoring and Management of Land and Water Resources in Agricultural Landscapes of Central Asia [Mueller, L., A. Saparov and 
G. Lischeid (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 61–73.
  1520. Pender, J., A. Mirzabaev, and E. Kato, 2009: Economic Analysis of Sustainable Land Management Options in Central Asia. Final Report for ADB. IFPRI/ICARDA.Washington, D.C., USA, 44 pages.
  1521. Pender, J., A. Mirzabaev, and E. Kato, 2009: Economic Analysis of Sustainable Land Management Options in Central Asia. Final Report for ADB. IFPRI/ICARDA.Washington, D.C., USA, 44 pages.
  1522. Gupta, R. et al., 2009: Research prospectus: A vision for sustainable land management research in Central Asia. ICARDA Central Asia Caucasus program. Sustainable Agriculture in Central Asia and the Caucasus Series No.1, CGIAR-PFU, Tashkent, Uzbekistan, 84 pp.
  1523. Suleimenov, M., Z. Kaskarbayev, K. Akshalov, and N. Yushchenko, 2014: Conservation agriculture for long-term soil productivity. In: Novel Measurement and Assessment Tools for Monitoring and Management of Land and Water Resources in Agricultural Landscapes of Central Asia [Mueller, L., A. Saparov and G. Lischeid (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 441–454.
  1524. Mirzabaev, A., M. Ahmed, J. Werner, J. Pender, and M. Louhaichi, 2016a: Rangelands of Central Asia: Challenges and opportunities. J. Arid Land, 8, 93–108, doi:10.1007/s40333-015-0057-5.
  1525. Mirzabaev, A., J. Goedecke, O. Dubovyk, U. Djanibekov, Q.B. Le, and A. Aw-Hassan, 2016b: Economics of land degradation in Central Asia. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 261–290.
  1526. Alimaev, I.I. et al., 2008: The impact of livestock grazing on soils and vegetation around settlements in Southeast Kazakhstan. In: The Socio-Economic Causes and Consequences of Desertification in Central Asia, [Behnke R. (ed.)]. Springer Netherlands, Dordrecht, Netherlands, pp. 81–112.
  1527. Issanova, G., and J. Abuduwaili, 2017: Relationship between storms and land degradation. In: Aeolian Process as Dust Storms in the Deserts of Central Asia and Kazakhstan [Issanova, G. and J. Abuduwaili (eds.)]. Springer, Singapore, pp. 71–86.
  1528. Hoegh-Guldberg, O., D. Jacob, M. Taylor, M. Bindi, S. Brown, I. Camilloni, A. Diedhiou, R. Djalante, K.L. Ebi, F. Engelbrecht, J.Guiot, Y. Hijioka, S. Mehrotra, A. Payne, S.I. Seneviratne, A. Thomas, R. Warren, and G. Zhou, 2018: Impacts of 1.5°C of Global Warming on Natural and Human Systems. In: Global Warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty [Masson-Delmotte, V., P. Zhai, 
H.-O. Pörtner, D. Roberts, J. Skea, P.R. Shukla, A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, J.B.R. Matthews, Y. Chen, X. Zhou, M.I.Gomis, E. Lonnoy, T.Maycock, M.Tignor, and T. Waterfield (eds.)]. In press. pp. 175–311.
  1529. Reyer, C.P.O., 2017: Climate change impacts in Central Asia and their implications for development. Reg. Environ. Chang., 17, 1639–1650, doi:10.1007/s10113-015-0893-z.
  1530. Reyer, C.P.O., 2017: Climate change impacts in Central Asia and their implications for development. Reg. Environ. Chang., 17, 1639–1650, doi:10.1007/s10113-015-0893-z.
  1531. Hijioka, Y., E. Lin, J.J. Pereira, R.T. Corlett, X. Cui, G.E. Insarov, R.D. Lasco, 
E. Lindgren, and A. Surjan, 2014: Asia. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Barros, V.R., C.B. Field, D.J. Dokken, 
M.D. Mastrandrea, K.J. Mach, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White (eds.)]. Cambridge University Press, New York, USA, pp. 1327–1370.
  1532. Bekchanov, M. et al., 2010: Pros and cons of adopting water-wise approaches in the lower reaches of the Amu Darya: A socio-economic view. Water, 2, 200–216, doi:10.3390/w2020200.
  1533. Kust, G., 1999: Desertification: principles of ecological assessment and mapping. MSU-RAS Institute of Soil Science, Moscow, Russia, 362 pp.
  1534. Abdullaev, I., M. Ul Hassan, and K. Jumaboev, 2007: Water saving and economic impacts of land leveling: The case study of cotton production in Tajikistan. Irrig. Drain. Syst., 21, 251–263, doi:10.1007/s10795-007-9034-2.
  1535. Gupta, R. et al., 2009: Research prospectus: A vision for sustainable land management research in Central Asia. ICARDA Central Asia Caucasus program. Sustainable Agriculture in Central Asia and the Caucasus Series No.1, CGIAR-PFU, Tashkent, Uzbekistan, 84 pp.
  1536. Kust, G., J. Mott, N. Jain, T. Sampath, and A. Armstrong, 2014: SLM oriented projects in Tajikistan: Experience and lessons learned. Planet@Risk, 2 (1).
  1537. Nurbekov, A. et al., 2016: Conservation agriculture for combating land degradation in Central Asia: A synthesis. AIMS Agric. Food, 1, 144–156, doi:10.3934/agrfood.2016.2.144.
  1538. Gupta, R. et al., 2009: Research prospectus: A vision for sustainable land management research in Central Asia. ICARDA Central Asia Caucasus program. Sustainable Agriculture in Central Asia and the Caucasus Series No.1, CGIAR-PFU, Tashkent, Uzbekistan, 84 pp.
  1539. Kienzler, K.M. et al., 2012: Conservation agriculture in Central Asia – What do we know and where do we go from here? F. Crop. Res., 132, doi:10.1016/j.fcr.2011.12.008.
  1540. Pulatov, A. et al., 2012: Introducing conservation agriculture on irrigated meadow alluvial soils (arenosols) in Khorezm, Uzbekistan. In: Cotton, Water, Salts and Soums [Martius, C., I. Rudenko, J.P.A. Lamers and P.L.G. Vlek (eds.)]. Springer Netherlands, Dordrecht, Netherlands, pp. 195–217.
  1541. Mirzabaev, A., M. Ahmed, J. Werner, J. Pender, and M. Louhaichi, 2016a: Rangelands of Central Asia: Challenges and opportunities. J. Arid Land, 8, 93–108, doi:10.1007/s40333-015-0057-5.
  1542. Mirzabaev, A., J. Goedecke, O. Dubovyk, U. Djanibekov, Q.B. Le, and A. Aw-Hassan, 2016b: Economics of land degradation in Central Asia. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 261–290.
  1543. Djanibekov, U., and A. Khamzina, 2016: Stochastic economic assessment of afforestation on marginal land in irrigated farming system. Environ. Resour. Econ., 63, 95–117, doi:10.1007/s10640-014-9843-3.
  1544. Djanibekov, U. et al., 2016: Adoption of sustainable land uses in post-Soviet Central Asia: The case for agroforestry. Sustainability, 8, 1030, doi:10.3390/su8101030.
  1545. Khamzina, A., J.P.A. Lamers, and P.L.G. Vlek, 2009: Nitrogen fixation by Elaeagnus angustifolia in the reclamation of degraded croplands of Central Asia. Tree Physiol., 29, 799–808, doi:10.1093/treephys/tpp017.
  1546. Khamzina, A., J.P.A. Lamers, and C. Martius, 2016: Above- and below-ground litter stocks and decay at a multi-species afforestation site on arid, saline soil. Nutr. Cycl. Agroecosystems, 104, 187–199, doi:10.1007/s10705-016-9766-1.
  1547. Dukhovny, V.A., V.I. Sokolov, and D.R. Ziganshina, 2013: Integrated water resources management in Central Asia, as a way of survival in conditions of water scarcity. Quat. Int., 311, 181–188, doi:10.1016/J.QUAINT.2013.07.003.
  1548. Kazbekov, J., I. Abdullaev, H. Manthrithilake, A. Qureshi, and K. Jumaboev, 2009: Evaluating planning and delivery performance of water user associations (WUAs) in Osh Province, Kyrgyzstan. Agric. Water Manag., 96, 1259–1267, doi:10.1016/J.AGWAT.2009.04.002.
  1549. Akinshina, N., A. Azizov, T. Karasyova, and E. Klose, 2016: On the issue of halophytes as energy plants in saline environment. Biomass and Bioenergy, 91, 306–311, doi:10.1016/J.BIOMBIOE.2016.05.034.
  1550. Qadir, M., A.D. Noble, A.S. Qureshi, R.K. Gupta, T. Yuldashev, and A. Karimov, 2009: Salt-induced land and water degradation in the Aral Sea Basin: A challenge to sustainable agriculture in Central Asia. Nat. Resour. Forum, 33, 134–149, doi:10.1111/j.1477-8947.2009.01217.x.
  1551. Toderich, K.N. et al., 2009: Phytogenic resources of halophytes of Central Asia and their role for rehabilitation of sandy desert degraded rangelands. L. Degrad. Dev., 20, 386–396, doi:10.1002/ldr.936.
  1552. Toderich, K.N. et al., 2008: New approaches for biosaline agriculture development, management and conservation of sandy desert ecosystems. In: Biosaline Agriculture and High Salinity Tolerance [Abdelly, C., M. Öztürk, M. Ashraf, C. Grignon (eds.)]. Springer, Birkhäuser Basel, Switzerland, pp. 247–264.
  1553. Breckle, S.-W., 2013: From Aral Sea to Aralkum: An Ecological Disaster or Halophytes’ Paradise. Springer, Berlin, Germany, pp. 351–398.
  1554. Aw-Hassan, A. et al., 2016: Economics of land degradation and improvement in Uzbekistan. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 651–682.
  1555. Bekchanov, M., C. Ringler, A. Bhaduri, and M. Jeuland, 2016: Optimizing irrigation efficiency improvements in the Aral Sea Basin. Water Resour. Econ., 13, 30–45, doi:10.1016/J.WRE.2015.08.003.
  1556. Bobojonov, I., J.P.A. Lamers, M. Bekchanov, N. Djanibekov, J. Franz-Vasdeki, 
J. Ruzimov, and C. Martius, 2013: Options and constraints for crop diversification: A case study in sustainable agriculture in Uzbekistan. Agroecol. Sustain. Food Syst., 37, 788–811, doi:10.1080/21683565.2013.775539.
  1557. Djanibekov, U., and A. Khamzina, 2016: Stochastic economic assessment of afforestation on marginal land in irrigated farming system. Environ. Resour. Econ., 63, 95–117, doi:10.1007/s10640-014-9843-3.
  1558. Djanibekov, U. et al., 2016: Adoption of sustainable land uses in post-Soviet Central Asia: The case for agroforestry. Sustainability, 8, 1030, doi:10.3390/su8101030.
  1559. Hamidov, A., K. Helming, and D. Balla, 2016: Impact of agricultural land use in Central Asia: A review. Agron. Sustain. Dev., 36, 6, doi:10.1007/s13593-015-0337-7.
  1560. Mirzabaev, A., M. Ahmed, J. Werner, J. Pender, and M. Louhaichi, 2016a: Rangelands of Central Asia: Challenges and opportunities. J. Arid Land, 8, 93–108, doi:10.1007/s40333-015-0057-5.
  1561. Mirzabaev, A., J. Goedecke, O. Dubovyk, U. Djanibekov, Q.B. Le, and A. Aw-Hassan, 2016b: Economics of land degradation in Central Asia. In: Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development [Nkonya, E., A. Mirzabaev and J. von Braun (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 261–290.
  1562. Yurtoglu, M.A., 2015: The afforestation and erosion control mobilization action plan in Turkey. In: Living Land [Griffiths, J. (ed.)]. UNCCD and Tudor Rose, Leicester, United Kingdom, pp. 37–39.
  1563. Çalişkan, S., and M. Boydak, 2017: Afforestation of arid and semi-arid ecosystems in Turkey. Turkish J. Agric. For., 41, 317–330, doi:10.3906/tar-1702-39.
  1564. Amiraslani, F., and D. Dragovich, 2011: Combating desertification in Iran over the last 50 years: An overview of changing approaches. J. Environ. Manage., 92, 1–13, doi:10.1016/j.jenvman.2010.08.012.
  1565. Goffner, D., H. Sinare, and L.J. Gordon, 2019: The Great Green Wall for the Sahara and the Sahel Initiative as an opportunity to enhance resilience in Sahelian landscapes and livelihoods. Reg. Environ. Chang., 19, 1417–1428, doi:10.1007/s10113-019-01481-z.
  1566. State Forestry Administration of China, 2015: A Bulletin of Desertification and Sandification State of China. http://www.documentcloud.org/documents/1237947-state-forestry-administration-desertification.html, State Forestry Administration of China, Beijing, China, 32 pp.
  1567. Wang, X., F. Yamauchi, K. Otsuka, and J. Huang, 2014: Wage Growth, Landholding, and Mechanization in Chinese Agriculture. Policy Research Working Paper No. WPS 7138,World Bank Group, Washington, DC, USA, 43 pp.
  1568. Wang, F., X. Pan, D. Wang, C. Shen, and Q. Lu, 2013: Combating desertification in China: Past, present and future. Land Use Policy, 31, 311–313, doi:10.1016/j.landusepol.2012.07.010.
  1569. Wang, K., and R.E. Dickinson, 2013: Contribution of solar radiation to decadal temperature variability over land. Proc. Natl. Acad. Sci. U.S.A., 110, 14877–14882, doi:10.1073/pnas.1311433110.
  1570. State Forestry Administration of China, 2015: A Bulletin of Desertification and Sandification State of China. http://www.documentcloud.org/documents/1237947-state-forestry-administration-desertification.html, State Forestry Administration of China, Beijing, China, 32 pp.
  1571. State Forestry Administration of China, 2015: A Bulletin of Desertification and Sandification State of China. http://www.documentcloud.org/documents/1237947-state-forestry-administration-desertification.html, State Forestry Administration of China, Beijing, China, 32 pp.
  1572. State Forestry Administration of China, 2015: A Bulletin of Desertification and Sandification State of China. http://www.documentcloud.org/documents/1237947-state-forestry-administration-desertification.html, State Forestry Administration of China, Beijing, China, 32 pp.
  1573. Wang, F., X. Pan, D. Wang, C. Shen, and Q. Lu, 2013: Combating desertification in China: Past, present and future. Land Use Policy, 31, 311–313, doi:10.1016/j.landusepol.2012.07.010.
  1574. Wang, K., and R.E. Dickinson, 2013: Contribution of solar radiation to decadal temperature variability over land. Proc. Natl. Acad. Sci. U.S.A., 110, 14877–14882, doi:10.1073/pnas.1311433110.
  1575. Wang, X., F. Yamauchi, K. Otsuka, and J. Huang, 2014: Wage Growth, Landholding, and Mechanization in Chinese Agriculture. Policy Research Working Paper No. WPS 7138,World Bank Group, Washington, DC, USA, 43 pp.
  1576. Yang, L., J. Wu, and P. Shen, 2013: Roles of science in institutional changes: The case of desertification control in China. Environ. Sci. Policy, 27, 32–54, doi:10.1016/j.envsci.2012.10.017.
  1577. Yang, Y., Z. Wang, J. Li, C. Gang, Y. Zhang, Y. Zhang, I. Odeh, and J. Qi, 2016: Comparative assessment of grassland degradation dynamics in response to climate variation and human activities in China, Mongolia, Pakistan and Uzbekistan from 2000 to 2013. J. Arid Environ., 135, 164–172, doi:10.1016/J.JARIDENV.2016.09.004.
  1578. Jiang, H., 2016: Taking down the ‘Great Green Wall’: The science and policy discourse of desertification and its control in China. In: The End of Desertification? Disputing Environmental Change in the Drylands [Behnke R. and M. Mortimore (eds.)]. Springer, Berlin, Germany, pp. 513–536.
  1579. Wang, X.M., C.X. Zhang, E. Hasi, and Z.B. Dong, 2010: Has the Three-Norths Forest Shelterbelt Program solved the desertification and dust storm problems in arid and semi-arid China? J. Arid Environ., 74, 13–22, doi:10.1016/J.JARIDENV.2009.08.001.
  1580. Feng, X. et al., 2016: Revegetation in China’s Loess Plateau is approaching sustainable water resource limits. Nat. Clim. Chang., 6, 1019–1022, doi:10.1038/nclimate3092.
  1581. Feng, Q., H. Ma, X. Jiang, X. Wang, and S. Cao, 2015: What has caused desertification in China? Sci. Rep., 5, 15998, doi:10.1038/srep15998.
  1582. Feng, Q., H. Ma, X. Jiang, X. Wang, and S. Cao, 2015: What has caused desertification in China? Sci. Rep., 5, 15998, doi:10.1038/srep15998.
  1583. Wang, X., Y. Yang, Z. Dong, and C. Zhang, 2009: Responses of dune activity and desertification in China to global warming in the twenty-first century. Glob. Planet. Change, 67, 167–185, doi:10.1016/j.gloplacha.2009.02.004.
  1584. Wang, X.M., C.X. Zhang, E. Hasi, and Z.B. Dong, 2010: Has the Three-Norths Forest Shelterbelt Program solved the desertification and dust storm problems in arid and semi-arid China? J. Arid Environ., 74, 13–22, doi:10.1016/J.JARIDENV.2009.08.001.
  1585. Wang, H., Y. Pan, Y. Chen, and Z. Ye, 2017a: Linear trend and abrupt changes of climate indices in the arid region of north-western China. Atmos. Res., 196, 108–118. doi.org/10.1016/j.atmosres.2017.06.008.
  1586. Wang, X., T. Hua, L. Lang, and W. Ma, 2017b: Spatial differences of aeolian desertification responses to climate in arid Asia. Glob. Planet. Change, 148, 22–28, doi:10.1016/j.gloplacha.2016.11.008.
  1587. Wang, Y., J. Zhang, S. Tong, and E. Guo, 2017c: Monitoring the trends of aeolian desertified lands based on time-series remote sensing data in the Horqin Sandy Land, China. Catena, 157, 286–298, doi:10.1016/j.catena.2017.05.030.
  1588. Zhang, Z., and D. Huisingh, 2018: Combating desertification in China: Monitoring, control, management and revegetation. J. Clean. Prod., 182, 765–775, doi:10.1016/J.JCLEPRO.2018.01.233.
  1589. Yang, L., J. Wu, and P. Shen, 2013: Roles of science in institutional changes: The case of desertification control in China. Environ. Sci. Policy, 27, 32–54, doi:10.1016/j.envsci.2012.10.017.
  1590. Yang, Y., Z. Wang, J. Li, C. Gang, Y. Zhang, Y. Zhang, I. Odeh, and J. Qi, 2016: Comparative assessment of grassland degradation dynamics in response to climate variation and human activities in China, Mongolia, Pakistan and Uzbekistan from 2000 to 2013. J. Arid Environ., 135, 164–172, doi:10.1016/J.JARIDENV.2016.09.004.
  1591. Belaaz, M., 2003: Le barrage vert en tant que patrimoine naturel national et moyen de lutte contre la désertification. Dans actes du XII°. Proceedings of the XII World Forestry Congress 2003, Quebec, Canada. 0301-B3.
  1592. Bensaid, S., 1995: Bilan critique du barrage vert en Algérie. Sci. Chang. planétaires/Sécheresse, 6, 247–255.
  1593. Belala, F. et al., 2018: Rainfall patterns of Algerian steppes and the impacts on natural vegetation in the 20th century. J. Arid Land, 10, 561–573, doi:10.1007/s40333-018-0095-x.
  1594. Achite, M., S. Ouillon, H. Quoc Viet, and C. Giay, 2016: Recent changes in climate, hydrology and sediment load in the Wadi Abd, Algeria (1970–2010). Hydrol. Earth Syst. Sci, 20, 1355–1372, doi:10.5194/hess-20-1355-2016.
  1595. Salamani, M., H.K. Hanifi, A. Hirche, and D. Nedjraoui, 2012: Évaluation de la sensibilitÉ À la dÉsertification en Algerie. Rev. Écol. (Terre Vie), 67, 71–84.
  1596. Houyou, Z., C.L. Bielders, H.A. Benhorma, A. Dellal, and A. Boutemdjet, 2016: Evidence of strong land degradation by wind erosion as a result of rainfed cropping in the Algerian steppe: A case study at Laghouat. L. Degrad. Dev., 27, 1788–1796, doi:10.1002/ldr.2295.
  1597. Merdas, S., B. Nouar, and F. Lakhdari, 2015: The Green Dam in Algeria as a tool to combat desertification. Planet@Risk, 3, 3–6.
  1598. Sacande, M., 2018: Action Against Desertification, Land Restoration. Food and Agriculture Organization of the United Nations, Rome, Italy.
  1599. Mbow, C., 2017: The Great Green Wall in the Sahel. Oxford Research Encylopedia, Oxford University Press, UK, doi: 10.1093/acrefore/9780190228620.013.559.
  1600. Benjaminsen, T.A. and P. Hiernaux, 2019: From desiccation to global climate change: A history of the desertification narrative in the West African Sahel, 1900-2018. Glob. Environ., 12, 206–236, doi:10.3197/ge.2019.120109.
  1601. Mbow, C., 2017: The Great Green Wall in the Sahel. Oxford Research Encylopedia, Oxford University Press, UK, doi: 10.1093/acrefore/9780190228620.013.559.
  1602. Sacande, M., 2018: Action Against Desertification, Land Restoration. Food and Agriculture Organization of the United Nations, Rome, Italy.
  1603. Bradley, B.A., D.M. Blumenthal, D.S. Wilcove, and L.H. Ziska, 2010: Predicting plant invasions in an era of global change. Trends Ecol. Evol., 25, 310–318, doi:10.1016/J.TREE.2009.12.003.
  1604. Davis, M.A., J.P. Grime, and K. Thompson, 2000: Fluctuating resources in plant communities: A general theory of invasibility. J. Ecol., 88, 528–534, doi:10.1046/j.1365-2745.2000.00473.x.
  1605. Bradley, B.A., D.M. Blumenthal, D.S. Wilcove, and L.H. Ziska, 2010: Predicting plant invasions in an era of global change. Trends Ecol. Evol., 25, 310–318, doi:10.1016/J.TREE.2009.12.003.
  1606. Settele, J., R. Scholes, R. Betts, S. Bunn, P. Leadley, D. Nepstad, J.T. Overpeck, and M.A. Taboada, 2014: Terrestrial and inland water systems. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, 
M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, 
A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White, (eds.)]. Cambridge University Press, New York, USA, pp. 271–359.
  1607. Scasta, J.D., D.M. Engle, S.D. Fuhlendorf, D.D. Redfearn, and T.G. Bidwell, 2015: Meta-analysis of exotic forages as invasive plants in complex multi-functioning landscapes. Invasive Plant Sci. Manag., 8, 292–306, doi:10.1614/ipsm-d-14-00076.1.
  1608. Ehrenfeld, J.G., 2003: Effects of Exotic plant invasions on soil nutrient cycling processes. Ecosystems, 6, 503–523, doi:10.1007/s10021-002-0151-3.
  1609. Bradley, B.A. et al., 2012: Global change, global trade, and the next wave of plant invasions. Front. Ecol. Environ., 10, 20–28, doi:10.1890/110145.
  1610. Newton, A.C., C. Echeverría, E. Cantarello, and G. Bolados, 2011: Projecting impacts of human disturbances to inform conservation planning and management in a Dryland forest landscape. Biol. Conserv., 144, 1949–1960, doi:10.1016/J.BIOCON.2011.03.026.
  1611. Safriel, U., Adeel, Z. et al., 2005: Dryland systems. In: Ecosystems and Human Well-Being: Current State and Trends [Hassan, R., R.J. Scholes, and N. Ash, (eds.)]. Island Press, Washington, DC, USA, pp. 623–662.
  1612. Chytrý, M. et al., 2012: Projecting trends in plant invasions in Europe under different scenarios of future land-use change. Glob. Ecol. Biogeogr., 21, 75–87, doi:10.1111/j.1466-8238.2010.00573.x.
  1613. Richardson, D.M. et al., 2011: Human-mediated introductions of Australian acacias – A global experiment in biogeography. Divers. Distrib., 17, 771–787, doi:10.1111/j.1472-4642.2011.00824.x.
  1614. Seebens, H. et al., 2018: Global rise in emerging alien species results from increased accessibility of new source pools. Proc. Natl. Acad. Sci. U.S.A., 115, E2264–E2273, doi:10.1073/pnas.1719429115.
  1615. Cassidy, T.M., J.H. Fownes, and R.A. Harrington, 2004: Nitrogen limits an invasive perennial shrub in forest understory. Biol. Invasions, 6, 113–121, doi:10.1023/B:BINV.0000010128.44332.0f.
  1616. Reynolds, J.F., P.R. Kemp, K. Ogle, and R.J. Fernández, 2004: Modifying the ‘pulse–reserve’ paradigm for deserts of North America: Precipitation pulses, soil water, and plant responses. Oecologia, 141, 194–210, doi:10.1007/s00442-004-1524-4.
  1617. Sala, A., D. Verdaguer, and M. Vila, 2006: Sensitivity of the invasive geophyte Oxalis pes-caprae to nutrient availability and competition. Ann. Bot., 99, 637–645, doi:10.1093/aob/mcl289.
  1618. Helldén, U., and C. Tottrup, 2008: Regional desertification: A global synthesis. Glob. Planet. Change, 64, 169–176, doi:10.1016/j.gloplacha.2008.10.006.
  1619. Hellmann, J.J., J.E. Byers, B.G. Bierwagen, and J.S. Dukes, 2008: Five potential consequences of climate change for invasive species. Conserv. Biol., 22, 534–543, doi:10.1111/j.1523-1739.2008.00951.x.
  1620. Bradley, B.A., D.M. Blumenthal, D.S. Wilcove, and L.H. Ziska, 2010: Predicting plant invasions in an era of global change. Trends Ecol. Evol., 25, 310–318, doi:10.1016/J.TREE.2009.12.003.
  1621. Eldridge, D.J. et al., 2011: Impacts of shrub encroachment on ecosystem structure and functioning: Towards a global synthesis. Ecol. Lett., 14, 709–722, doi:10.1111/j.1461-0248.2011.01630.x.
  1622. Rouget, M., M.P. Robertson, J.R.U. Wilson, C. Hui, F. Essl, J.L. Renteria, and 
D.M. Richardson, 2016: Invasion debt – Quantifying future biological invasions. Divers. Distrib., 22, 445–456, doi:10.1111/ddi.12408.
  1623. Bellard, C., W. Thuiller, B. Leroy, P. Genovesi, M. Bakkenes, and F. Courchamp, 2013: Will climate change promote future invasions? Glob. Chang. Biol., 19, 3740–3748, doi:10.1111/gcb.12344.
  1624. Helldén, U., and C. Tottrup, 2008: Regional desertification: A global synthesis. Glob. Planet. Change, 64, 169–176, doi:10.1016/j.gloplacha.2008.10.006.
  1625. Hellmann, J.J., J.E. Byers, B.G. Bierwagen, and J.S. Dukes, 2008: Five potential consequences of climate change for invasive species. Conserv. Biol., 22, 534–543, doi:10.1111/j.1523-1739.2008.00951.x.
  1626. Seebens, H. et al., 2015: Global trade will accelerate plant invasions in emerging economies under climate change. Glob. Chang. Biol., 21, 4128–4140, doi:10.1111/gcb.13021.
  1627. Brooks, M.L. et al., 2004: Effects of invasive alien plants on fire regimes. Bioscience, 54, 677–688, doi:10.1641/0006-3568(2004)054[0677:eoiapo]2.0.co;2.
  1628. Grace, J.B., M.D. Smith, S.L. Grace, S.L.C., and T.J. Stohlgren, 2001: Interactions between fire and invasive plants in temperate grasslands of North America. In: Proceedings of the Invasive Species Workshop: The Role of Fire in the Control and Spread of Invasive Species [Galley, K.E.M. and T.P. Wilson (eds.)]. Tall Timbers Research Station, Florida, USA, pp. 40–65.
  1629. Keeley, J.E., and T.J. Brennan, 2012: Fire-driven alien invasion in a fire-adapted ecosystem. Oecologia, 169, 1043–1052, doi:10.1007/s00442-012-2253-8.
  1630. Adkins, S. and A. Shabbir, 2014: Biology, ecology and management of the invasive parthenium weed (Parthenium hysterophorus L.). Pest Manag. Sci., 70, 1023–1029, doi:10.1002/ps.3708.
  1631. Teka, K., 2016: Parthenium hysterophorus (asteraceae) expansion, environmental impact and controlling strategies in Tigray, Northern Ethiopia: A review. J. Drylands, 6, 434–448.
  1632. Tamado, T., W. Schutz, and P. Milberg, 2002: Germination ecology of the weed Parthenium hysterophorus in eastern Ethiopia. Ann. Appl. Biol., 140, 263–270, doi:10.1111/j.1744-7348.2002.tb00180.x.
  1633. Mekonnen, G., 2017: Threats and management options of Parthenium (Parthenium hysterophorus l.) in Ethiopia. Agric. Res. Technol., 10, 1–7, doi:10.19080/ARTOAJ.2017.10.555798.
  1634. Patel, S., 2011: Harmful and beneficial aspects of Parthenium hysterophorus: An update. 3 Biotech, 1, 1–9, doi:10.1007/s13205-011-0007-7.
  1635. Mainali, K.P. et al., 2015: Projecting future expansion of invasive species: Comparing and improving methodologies for species distribution modeling. Glob. Chang. Biol., 21, 4464–4480, doi:10.1111/gcb.13038.
  1636. Fasil, R., 2011: Parthenium Hysterophorus in Ethiopia: Distribution and Importance, and Current Efforts to Manage the Scourge. Proceedings of Workshop on Noxious Weeds in Production of Certified Seeds, 11–12 July. Food and Agriculture Organization of the United Nations, Accra, Ghana.
  1637. Castellanos-Villegas, A.E., G. Yanes, and D. Valdes, 2002: Drought-tolerant exotic buffel-grass and desertification. In: Proceedings of a North American Conference Weeds Across Borders Arizona-Sonora Desert Museum [Tellman, B. (ed.)]. Arizona-Sonora Desert Museum, Tucson, USA. pp. 99–112.
  1638. Arriaga, L., A.E. Castellanos, E. Moreno, and J. Alarcón, 2004: Potential ecological distribution of alien invasive species and risk assessment: A case study of buffel grass in arid regions of Mexico. Conserv. Biol., 18, 1504–1514, doi:10.1111/j.1523-1739.2004.00166.x.
  1639. Marshall, M., C. Funk, and J. Michaelsen, 2012a: Examining evapotranspiration trends in Africa. Clim. Dyn., 38, 1849–1865, doi:10.1007/s00382-012-1299-y.
  1640. Marshall, V.M., M.M. Lewis, and B. Ostendorf, 2012b: Buffel grass (Cenchrus ciliaris) as an invader and threat to biodiversity in arid environments: 
A review. J. Arid Environ., 78, 1–12, doi:10.1016/J.JARIDENV.2011.11.005.
  1641. Miller, G., M. Friedel, P. Adam, and V. Chewings, 2010: Ecological impacts of buffel grass (Cenchrus ciliaris L.) invasion in Central Australia – Does field evidence support a fire-invasion feedback? Rangel. J., 32, 353, doi:10.1071/RJ09076.
  1642. Schlesinger, C., S. White, and S. Muldoon, 2013: Spatial pattern and severity of fire in areas with and without buffel grass ( Cenchrus ciliaris) and effects on native vegetation in central Australia. Austral Ecol., 38, 831–840, doi:10.1111/aec.12039.
  1643. Franklin, K., and F. Molina-Freaner, 2010: Consequences of buffelgrass pasture development for primary productivity, perennial plant richness, and vegetation structure in the drylands of Sonora, Mexico. Conserv. Biol., 24, 1664–1673, doi:10.1111/j.1523-1739.2010.01540.x.
  1644. Miller, R.F. et al., 2011: Characteristics of sagebrush habitats and limitations to long-term conservation. Greater sage-grouse: Ecology and conservation of a landscape species and its habitats. Stud. Avian Biol., 38, 145–184.
  1645. Balch, J.K., B.A. Bradley, C.M. D’Antonio, and J. Gómez-Dans, 2013: Introduced annual grass increases regional fire activity across the arid western USA (1980–2009). Glob. Chang. Biol., 19, 173–183, doi:10.1111/gcb.12046.
  1646. Balch, J.K., B.A. Bradley, C.M. D’Antonio, and J. Gómez-Dans, 2013: Introduced annual grass increases regional fire activity across the arid western USA (1980–2009). Glob. Chang. Biol., 19, 173–183, doi:10.1111/gcb.12046.
  1647. Balch, J.K., B.A. Bradley, C.M. D’Antonio, and J. Gómez-Dans, 2013: Introduced annual grass increases regional fire activity across the arid western USA (1980–2009). Glob. Chang. Biol., 19, 173–183, doi:10.1111/gcb.12046.
  1648. Pilliod, D.S., J.L. Welty, and R.S. Arkle, 2017: Refining the cheatgrass-fire cycle in the Great Basin: Precipitation timing and fine fuel composition predict wildfire trends. Ecol. Evol., 7, 8126–8151, doi:10.1002/ece3.3414.
  1649. Crawford, J.A., Olson, R.A., West, N.E., Mosley, J.C., Schroeder, 
M.A., Whitson, T.D., Miller, R.F., Gregg, M.A., Boyd, C.S., 2004: Ecology and management of sage-grouse and sage-grouse habitat. J. Range Manag., 57, 2–19, doi:10.2111/1551-5028(2004)057[0002:EAMOSA]2.0.CO;2.
  1650. Larrucea, E.S., and P.F. Brussard, 2008: Habitat selection and current distribution of the pygmy rabbit in Nevada and California, USA,.J. Mammal., 89, 691–699, doi:10.1644/07-MAMM-A-199R.1.
  1651. Lockyer, Z.B., P.S. Coates, M.L. Casazza, S. Espinosa, and D.J. Delehanty, 2015: Nest-site selection and reproductive success of greater sage-grouse in a fire-affected habitat of north-western Nevada. J. Wildl. Manage., 79, 785–797, doi:10.1002/jwmg.899.
  1652. Hull, A.C. and G. Stewart, 1949: Replacing Cheatgrass by Reseeding with Perennial Grass on Southern Idaho Ranges. American Society of Agronomy Journal, 40, 694–703.
  1653. Davies, G.M. et al., 2012: Trajectories of change in sagebrush steppe vegetation communities in relation to multiple wildfires. Ecol. Appl., 22, 1562–1577, doi:10.1890/10-2089.1.
  1654. Taylor, K., T. Brummer, L.J. Rew, M. Lavin, and B.D. Maxwell, 2014: Bromus tectorum response to fire varies with climate conditions. Ecosystems, 17, 960–973, doi:10.1007/s10021-014-9771-7.
  1655. Bureau of Land Management, 2005: Vegetation Treatments on Bureau of Land Management Lands in 17 Western State. Nevada State Office, Reno, Nevada, USA. 497 pp.
  1656. Arkle, R.S. et al., 2014: Quantifying restoration effectiveness using multi-scale habitat models: Implications for sage-grouse in the Great Basin. Ecosphere, 5, art31, doi:10.1890/ES13-00278.1.
  1657. Knutson, K.C. et al., 2014: Long-term effects of seeding after wildfire on vegetation in Great Basin shrubland ecosystems. J. Appl. Ecol., 51, 1414–1424, doi:10.1111/1365-2664.12309.
  1658. Chambers, J.C. et al., 2014: Resilience to stress and disturbance, and resistance to Bromus tectorum L. Invasion in cold desert shrublands of western North America. Ecosystems, 17, 360–375, doi:10.1007/s10021-013-9725-5.
  1659. Miller, R.F. et al., 2011: Characteristics of sagebrush habitats and limitations to long-term conservation. Greater sage-grouse: Ecology and conservation of a landscape species and its habitats. Stud. Avian Biol., 38, 145–184.
  1660. Pellant, M., B. Abbey, and S. Karl, 2004: Restoring the Great Basin Desert, USA: Integrating science, management, and people. Environ. Monit. Assess., 99, 169–179, doi:10.1007/s10661-004-4017-3.
  1661. Hernandez, R.R., and D.R. Sandquist, 2011: Disturbance of biological soil crust increases emergence of exotic vascular plants in California sage scrub. Plant Ecol., 212, 1709–1721, doi:10.1007/s11258-011-9943-x.
  1662. Abatzoglou, J.T. and C.A. Kolden, 2011: Climate change in western US deserts: Potential for increased wildfire and invasive annual grasses. Rangel. Ecol. Manag., 64, 471–478, doi:10.2111/REM-D-09-00151.1.
  1663. Harpold, A.A., and P.D. Brooks, 2018: Humidity determines snowpack ablation under a warming climate. Proc. Natl. Acad. Sci. U.S.A., 115, 1215–1220, doi:10.1073/pnas.1716789115.
  1664. Mote, P.W., A.F. Hamlet, M.P. Clark, and D.P. Lettenmaier, 2005: Declining mountain snowpack in western North America. Bull. Am. Meteorol. Soc., 86, 39–49, doi:10.1175/BAMS-86-1-39.
  1665. Khan, M.A. et al., 2010: Invasive species of federal capital area Islamabad, Pakistan. Pakistan J. Bot., 42, 1529–1534.
  1666. Qureshi, H., M. Arshad, and Y. Bibi, 2014: Invasive flora of Pakistan: A critical analysis. Int. J. Biosci., 6655, 407–424, doi:10.12692/ijb/4.1.407-424.
  1667. Marwat, K.B., S. Hashim, and H. Ali, 2010: Weed management: A case study from north-west Pakistan. Pakistan J. Bot., 42, 341–353.
  1668. Rashid, M., S.H. Abbas, and A. Rehman, 2014: The Status of highly alien invasive plants in Pakistan and their impact on the ecosystem: A Review. Innovare J. Agric. Sci,. 2, 2–5.
  1669. Rashid, M., S.H. Abbas, and A. Rehman, 2014: The Status of highly alien invasive plants in Pakistan and their impact on the ecosystem: A Review. Innovare J. Agric. Sci,. 2, 2–5.
  1670. Rashid, M., S.H. Abbas, and A. Rehman, 2014: The Status of highly alien invasive plants in Pakistan and their impact on the ecosystem: A Review. Innovare J. Agric. Sci,. 2, 2–5.
  1671. Marwat, K.B., S. Hashim, and H. Ali, 2010: Weed management: A case study from north-west Pakistan. Pakistan J. Bot., 42, 341–353.
  1672. Rashid, M., S.H. Abbas, and A. Rehman, 2014: The Status of highly alien invasive plants in Pakistan and their impact on the ecosystem: A Review. Innovare J. Agric. Sci,. 2, 2–5.
  1673. Qureshi, H., M. Arshad, and Y. Bibi, 2014: Invasive flora of Pakistan: A critical analysis. Int. J. Biosci., 6655, 407–424, doi:10.12692/ijb/4.1.407-424.
  1674. Rashid, M., S.H. Abbas, and A. Rehman, 2014: The Status of highly alien invasive plants in Pakistan and their impact on the ecosystem: A Review. Innovare J. Agric. Sci,. 2, 2–5.
  1675. Jones, M.D. et al., 2019: 20,000 years of societal vulnerability and adaptation to climate change in Southwest Asia. Wiley Interdiscip. Rev. Water, 6, e1330, doi:10.1002/wat2.1330.
  1676. Jones, M.D. et al., 2019: 20,000 years of societal vulnerability and adaptation to climate change in Southwest Asia. Wiley Interdiscip. Rev. Water, 6, e1330, doi:10.1002/wat2.1330.
  1677. Abatzoglou, J.T. and C.A. Kolden, 2011: Climate change in western US deserts: Potential for increased wildfire and invasive annual grasses. Rangel. Ecol. Manag., 64, 471–478, doi:10.2111/REM-D-09-00151.1.
  1678. Ashkenazy, Y., H. Yizhaq, and H. Tsoar, 2012: Sand dune mobility under climate change in the Kalahari and Australian deserts. Clim. Change, 112, 901–923, doi:10.1007/s10584-011-0264-9.
  1679. Bachelet, D., K. Ferschweiler, T. Sheehan, and J. Strittholt, 2016: Climate change effects on southern California deserts. J. Arid Environ., 127, 17–29, doi:10.1016/J.JARIDENV.2015.10.003.
  1680. Guan, C., X. Li, P. Zhang, and C. Li, 2018: Effect of global warming on soil respiration and cumulative carbon release in biocrust-dominated areas in the Tengger Desert, northern China. J. Soils Sediments, 19, 1161–1170, doi:10.1007/s11368-018-2171-4.
  1681. Iknayan, K.J., and S.R. Beissinger, 2018: Collapse of a desert bird community over the past century driven by climate change. Proc. Natl. Acad. Sci. U.S.A., 115, 8597–8602, doi:10.1073/pnas.1805123115.
  1682. Ling, H., H. Xu, J. Fu, Z. Fan, and X. Xu, 2013: Suitable oasis scale in a typical continental river basin in an arid region of China: A case study of the Manas River Basin. Quat. Int., 286, 116–125, doi:10.1016/j.quaint.2012.07.027.
  1683. Ministry of Agriculture and Water Resources of Tunisia, and GIZ, 2007: Stratégie nationale d’adaptation de l’agriculture tunisienne et des écosystèmes aux changements climatiques.Eschborn, Germany, GIZ, 51 pages.
  1684. Jilali, A., 2014: Impact of climate change on the Figuig aquifer using 
a numerical model: Oasis of Eastern Morocco. J. Biol. Earth Sci., 4, 16–24.
  1685. Karmaoui, A., I. Ifaadassan, A. Babqiqi, M. Messouli, and Y.M. Khebiza, 2016: Analysis of the water supply-demand relationship in the Middle Draa Valley, Morocco, under climate change and socio-economic scenarios. J. Sci. Res. Reports, 9, 1–10.
  1686. Chowdhury, S., and M. Al-Zahrani, 2013: Implications of climate change on water resources in Saudi Arabia. Arab. J. Sci. Eng., 38, 1959–1971, doi:10.1007/s13369-013-0565-6.
  1687. Tarawneh, Q., and S. Chowdhury, 2018: Trends of climate change in Saudi Arabia: Implications on water resources. Climate, 6, 8, doi:10.3390/cli6010008.
  1688. Almazroui, M., M.N. Islam, K.S. Balkhair, Z. Şen, and A. Masood, 2017: Rainwater harvesting possibility under climate change: A basin-scale case study over western province of Saudi Arabia. Atmos. Res., 189, 11–23, doi:10.1016/J.ATMOSRES.2017.01.004.
  1689. Aldababseh, A. et al., 2018: Multi-criteria evaluation of irrigated agriculture suitability to achieve food security in an arid environment. Sustainability, 10, 803, doi:10.3390/su10030803.
  1690. Shabani, F., L. Kumar, and S. Taylor, 2015: Distribution of date palms in the Middle East based on future climate scenarios. Exp. Agric., 51, 244–263, doi:10.1017/S001447971400026X.
  1691. Safronov, O. et al., 2017: Detecting early signs of heat and drought stress in Phoenix dactylifera (date palm). PloS One, 12, e0177883, doi:10.1371/journal.pone.0177883.
  1692. Aldababseh, A. et al., 2018: Multi-criteria evaluation of irrigated agriculture suitability to achieve food security in an arid environment. Sustainability, 10, 803, doi:10.3390/su10030803.
  1693. Shabani, F., L. Kumar, and S. Taylor, 2015: Distribution of date palms in the Middle East based on future climate scenarios. Exp. Agric., 51, 244–263, doi:10.1017/S001447971400026X.
  1694. Al-Kalbani, M.S., M.F. Price, T. O’Higgins, M. Ahmed, and A. Abahussain, 2016: Integrated environmental assessment to explore water resources management in Al Jabal Al Akhdar, Sultanate of Oman. Reg. Environ. Chang., 16, 1345–1361, doi:10.1007/s10113-015-0864-4.
  1695. Luedeling, E., J. Gebauer, and A. Buerkert, 2009: Climate change effects on winter chill for tree crops with chilling requirements on the Arabian Peninsula. Clim. Change, 96, 219–237, doi:10.1007/s10584-009-9581-7.
  1696. Al-Kalbani, M. et al., 2014: Vulnerability assessment of environmental and climate change impacts on water resources in Al Jabal Al Akhdar, Sultanate of Oman. Water, 6, 3118–3135, doi:10.3390/w6103118.
  1697. Aoun-Sebaiti, B., A. Hani, L. Djabri, H. Chaffai, I. Aichouri, and N. Boughrira, 2014: Simulation of water supply and water demand in the valley of Seybouse (East Algeria). Desalin. Water Treat., 52, 2114–2119, doi:10.1080/19443994.2013.855662.
  1698. Switzman, H., B. Salem, M. Gad, Z. Adeel, and P. Coulibaly, 2018: Conservation planning as an adaptive strategy for climate change and groundwater depletion in Wadi El Natrun, Egypt. Hydrogeol. J., 26, 689–703, doi:10.1007/s10040-017-1669-y.
  1699. Johannsen, I. et al., 2016: Future of water supply and demand in the middle Drâa Valley, Morocco, under climate and land use change. Water, 8, 313, doi:10.3390/w8080313.
  1700. Karmaoui, A., I. Ifaadassan, A. Babqiqi, M. Messouli, and Y.M. Khebiza, 2016: Analysis of the water supply-demand relationship in the Middle Draa Valley, Morocco, under climate change and socio-economic scenarios. J. Sci. Res. Reports, 9, 1–10.
  1701. Sefelnasr, A., W. Gossel, and P. Wycisk, 2014: Three-dimensional groundwater flow modeling approach for the groundwater management options for the Dakhla Oasis, Western Desert, Egypt. Environ. Earth Sci., 72, 1227–1241, doi:10.1007/s12665-013-3041-4.
  1702. Siebert, C. et al., 2016: New tools for coherent information base for IWRM in arid regions: The upper mega aquifer system on the Arabian peninsula. In: Integrated Water Resources Management: Concept, Research and Implementation [Borchardt, D., J.J. Bogardi and R.B. Ibisch (eds.)]. Springer International Publishing, Cham, Switzerland, pp. 85–106.
  1703. Besser, H. et al., 2017: GIS-based evaluation of groundwater quality and estimation of soil salinization and land degradation risks in an arid Mediterranean site (SW Tunisia). Arab. J. Geosci., 10, 350, doi:10.1007/s12517-017-3148-0.
  1704. Ben Hassine, H., A. Ben Slimane, M. Mlawah, L. Albouchi, and A. Gandouzi, 2013: Effects of underground water on soil salinity and dates production in Kebili Oases Area (Tunisia): The case of El Bahaier Oasis. IOSR J. Environ. Sci. Toxicol. Food Technol., 4, 51–58, doi:10.9790/2402-0445158.
  1705. Abo-Ragab, S., 2010: A desertification impact on Siwa Oasis: Present and future challenges. Res. J. Agric. Biol. Sci., 6, 791–805. INSInet Publication.
  1706. Masoud, A.A. and K. Koike, 2006: Arid land salinization detected by remotely-sensed landcover changes: A case study in the Siwa region, NW Egypt. J. Arid Environ., 66, 151–167, doi:10.1016/J.JARIDENV.2005.10.011.
  1707. Abo-Ragab, S., Zaghloul, A.Q., 2017: Sand Dunes Movement and its Impact on Development Plans, Siwa Oasis. International Journal of Research in Economics and Social Sciences, 5:11-124-133, ISSN 2249-7382.
  1708. Masoud, A.A. and K. Koike, 2006: Arid land salinization detected by remotely-sensed landcover changes: A case study in the Siwa region, NW Egypt. J. Arid Environ., 66, 151–167, doi:10.1016/J.JARIDENV.2005.10.011.
  1709. Abo-Ragab, S., 2010: A desertification impact on Siwa Oasis: Present and future challenges. Res. J. Agric. Biol. Sci., 6, 791–805. INSInet Publication.
  1710. Askri, B., R. Bouhlila, and J.O. Job, 2010: Development and application of a conceptual hydrologic model to predict soil salinity within modern Tunisian oases. J. Hydrol., 380, 45–61, doi:10.1016/J.JHYDROL.2009.10.022.
  1711. Marlet, S., F. Bouksila, and A. Bahri, 2009: Water and salt balance at irrigation scheme scale: A comprehensive approach for salinity assessment in a Saharan oasis. Agric. Water Manag., 96, 1311–1322, doi:10.1016/J.AGWAT.2009.04.016.
  1712. Aldababseh, A. et al., 2018: Multi-criteria evaluation of irrigated agriculture suitability to achieve food security in an arid environment. Sustainability, 10, 803, doi:10.3390/su10030803.
  1713. Sefelnasr, A., W. Gossel, and P. Wycisk, 2014: Three-dimensional groundwater flow modeling approach for the groundwater management options for the Dakhla Oasis, Western Desert, Egypt. Environ. Earth Sci., 72, 1227–1241, doi:10.1007/s12665-013-3041-4.
  1714. Switzman, H., B. Salem, M. Gad, Z. Adeel, and P. Coulibaly, 2018: Conservation planning as an adaptive strategy for climate change and groundwater depletion in Wadi El Natrun, Egypt. Hydrogeol. J., 26, 689–703, doi:10.1007/s10040-017-1669-y.
  1715. Johannsen, I. et al., 2016: Future of water supply and demand in the middle Drâa Valley, Morocco, under climate and land use change. Water, 8, 313, doi:10.3390/w8080313.
  1716. Odhiambo, G.O., 2017: Water scarcity in the Arabian Peninsula and socio-economic implications. Appl. Water Sci., 7, 2479–2492, doi:10.1007/s13201-016-0440-1.
  1717. King, C., and D.S.G. Thomas, 2014: Monitoring environmental change and degradation in the irrigated oases of the Northern Sahara. J. Arid Environ., 103, 36–45, doi:10.1016/J.JARIDENV.2013.12.009.
  1718. Aldababseh, A. et al., 2018: Multi-criteria evaluation of irrigated agriculture suitability to achieve food security in an arid environment. Sustainability, 10, 803, doi:10.3390/su10030803.
  1719. Chen, J., S. Yin, H. Gebhardt, and X. Yang, 2018: Farmers’ livelihood adaptation to environmental change in an arid region: A case study of the Minqin Oasis, north-western China. Ecol. Indic., 93, 411–423, doi:10.1016/J.ECOLIND.2018.05.017.
  1720. Chen, T. et al., 2019: Disentangling the relative impacts of climate change and human activities on arid and semi-arid grasslands in Central Asia during 1982–2015. Sci. Total Environ., 653, 1311–1325, doi:10.1016/j.scitotenv.2018.11.058.
  1721. Chen, J., S. Yin, H. Gebhardt, and X. Yang, 2018: Farmers’ livelihood adaptation to environmental change in an arid region: A case study of the Minqin Oasis, north-western China. Ecol. Indic., 93, 411–423, doi:10.1016/J.ECOLIND.2018.05.017.
  1722. Chen, T. et al., 2019: Disentangling the relative impacts of climate change and human activities on arid and semi-arid grasslands in Central Asia during 1982–2015. Sci. Total Environ., 653, 1311–1325, doi:10.1016/j.scitotenv.2018.11.058.
  1723. Wang, H., Y. Pan, Y. Chen, and Z. Ye, 2017a: Linear trend and abrupt changes of climate indices in the arid region of north-western China. Atmos. Res., 196, 108–118. doi.org/10.1016/j.atmosres.2017.06.008.
  1724. Wang, X., T. Hua, L. Lang, and W. Ma, 2017b: Spatial differences of aeolian desertification responses to climate in arid Asia. Glob. Planet. Change, 148, 22–28, doi:10.1016/j.gloplacha.2016.11.008.
  1725. Wang, Y., J. Zhang, S. Tong, and E. Guo, 2017c: Monitoring the trends of aeolian desertified lands based on time-series remote sensing data in the Horqin Sandy Land, China. Catena, 157, 286–298, doi:10.1016/j.catena.2017.05.030.
  1726. German, L., H. Mansoor, G. Alemu, W. Mazengia, T. Amede, and A. Stroud, 2007: Participatory integrated watershed management: Evolution of concepts and methods in an ecoregional program of the eastern African highlands. Agric. Syst., 94, 189–204, doi:10.1016/J.AGSY.2006.08.008.
  1727. Al-Tabini, R., K. Al-Khalidi, and M. Al-Shudiefat, 2012: Livestock, medicinal plants and rangeland viability in Jordan’s Badia: Through the lens of traditional and local knowledge. Pastor. Res. Policy Pract., 2, 4, doi:10.1186/2041-7136-2-4.
  1728. AlSarmi, S., and R. Washington, 2011: Recent observed climate change over the Arabian Peninsula. J. Geophys. Res., 116, D11109, doi:10.1029/2010JD015459.
  1729. Tanarhte, M., P. Hadjinicolaou, and J. Lelieveld, 2015: Heatwave characteristics in the eastern Mediterranean and Middle East using extreme value theory. Clim. Res., 63, 99–113, doi:10.3354/cr01285.
  1730. De Pauw, E., M. Saba, and S. Ali, 2015: Mapping Climate Change in Iraq and Jordan. ICARDA Working Paper No. 27, Beirut, Lebanon, 141 pp., doi:10.13140/RG.2.1.1713.1365.
  1731. Giannakopoulos, C. et al., 2009: Climatic changes and associated impacts in the Mediterranean resulting from a 2°C global warming. Glob. Planet. Change, 68, 209–224.
  1732. Abu-Zanat, M.W., G.B. Ruyle, and N.F. Abdel-Hamid, 2004: Increasing range production from fodder shrubs in low rainfall areas. J. Arid Environ., 59, 205–216, doi:10.1016/J.JARIDENV.2003.12.011.
  1733. Oweis, T.Y., 2017: Rainwater harvesting for restoring degraded dry agropastoral ecosystems: A conceptual review of opportunities and constraints in a changing climate. Environ. Rev., 25, 135–149, doi:10.1139/er-2016-0069.
  1734. Tabieh, M. et al., 2015: Economic analysis of micro-catchment rainwater harvesting techniques in Jordan’s arid zones. Int. J. Appl. Environ. Sci., 10, 1205–1225.
  1735. Antinori, P., and V. Vallerani, 1994: Experiments in water harvesting technology with the dolphin and train ploughs. In: Water harvesting for improved agricultural production, [FAO, (ed.)]. Food and Agriculture Organization of the United Nations, Rome, Italy, pp. 113–132.
  1736. Gammoh, I.A., and T.Y. Oweis, 2011: Performance and adaptation of the Vallerani mechanized water harvesting system in degraded Badia Rangelands. J. Environ. Sci. Eng., 5, 1370–1380.
  1737. Ngigi, S.N., 2003: What is the limit of up-scaling rainwater harvesting in a river basin? Phys. Chem. Earth, Parts A/B/C, 28, 943–956, doi:10.1016/J.PCE.2003.08.015.
  1738. Oweis, T.Y., 2017: Rainwater harvesting for restoring degraded dry agropastoral ecosystems: A conceptual review of opportunities and constraints in a changing climate. Environ. Rev., 25, 135–149, doi:10.1139/er-2016-0069.
  1739. Oweis, T.Y., 2017: Rainwater harvesting for restoring degraded dry agropastoral ecosystems: A conceptual review of opportunities and constraints in a changing climate. Environ. Rev., 25, 135–149, doi:10.1139/er-2016-0069.
  1740. Tabieh, M. et al., 2015: Economic analysis of micro-catchment rainwater harvesting techniques in Jordan’s arid zones. Int. J. Appl. Environ. Sci., 10, 1205–1225.
  1741. Fried, T., L. Weissbrod, Y. Tepper, and G. Bar-Oz, 2018: A glimpse of an ancient agricultural ecosystem based on remains of micromammals in the Byzantine Negev Desert. R. Soc. Open Sci., 5, 171528, doi:10.1098/rsos.171528.
  1742. Stavi, I. et al., 2017: Ancient to recent-past runoff harvesting agriculture in recharge playas of the hyper-arid southern Israel. Water, 9, 991, doi:10.3390/w9120991.
  1743. Beckers, B., J. Berking, and B. Schütt, 2013: Ancient water harvesting methods in the drylands of the Mediterranean and Western Asia. eTopoi. J. Anc. Stud., 2, 145–164.
  1744. Bouma, J., D. Van Soest, and E. Bulte, 2007: How sustainable is participatory watershed development in India? Agric. Econ., 36, 13–22, doi:10.1111/j.1574-0862.2007.00173.x.
  1745. Kerr, J.M., G. Pangare, and V. Pangare, 2002: Watershed development projects in India: An evaluation. Research Report 127, International Food Policy Research Institute, Washington, DC, USA.
  1746. Bhalla, R.S., K.V. Devi Prasad, and N.W. Pelkey, 2013: Impact of India’s watershed development programs on biomass productivity. Water Resour. Res., 49, 1568–1580, doi:10.1002/wrcr.20133.
  1747. Ratna Reddy, V., M. Gopinath Reddy, S. Galab, J. Soussan, and 
O. Springate-Baginski, 2004: Participatory watershed development in India: Can it sustain rural livelihoods? Dev. Change, 35, 297–326, doi:10.1111/j.1467-7660.2004.00353.x.
  1748. Bhalla, R.S., K.V. Devi Prasad, and N.W. Pelkey, 2013: Impact of India’s watershed development programs on biomass productivity. Water Resour. Res., 49, 1568–1580, doi:10.1002/wrcr.20133.
  1749. Kerr, J.M., G. Pangare, and V. Pangare, 2002: Watershed development projects in India: An evaluation. Research Report 127, International Food Policy Research Institute, Washington, DC, USA.
  1750. Samuel, A. et al., 2007: Watershed Development in Maharashtra: Present Scenario and Issues for Restructuring the Programme. http://www.indiawater
portal.org/articles/watershed-development-maharashtra-present-scenario-
and-issues-restructuring-programme. Society for Promoting Participative 
Ecosystem Management (SOPPECOM), Pune, India, 158 pp.
  1751. Batchelor, C.H., M.S. Rama Mohan Rao, and S. Manohar Rao, 2003: Watershed development: A solution to water shortages in semi-arid India or part of the problem? L. Use Water Resour. Res., 3, 10, doi:10.22004/ag.econ.47866.
  1752. Glendenning, C.J., F.F. Van Ogtrop, A.K. Mishra, and R.W. Vervoort, 2012: Balancing watershed and local scale impacts of rainwater harvesting in India – A review. Agric. Water Manag., 107, 1–13.
  1753. Bharucha, Z.P., D. Smith, and J. Pretty, 2014: All paths lead to rain: Explaining why watershed development in India does not alleviate the experience of water scarcity. J. Dev. Stud., 50, 1209–1225, doi:10.1080/00220388.2014.
928699.
  1754. Kerr, J.M., G. Pangare, and V. Pangare, 2002: Watershed development projects in India: An evaluation. Research Report 127, International Food Policy Research Institute, Washington, DC, USA.
  1755. Vaidyanathan, A., 2006: Restructuring watershed development programmes.
Economic & Political Weekly, 41, 2984–2987. http://www.epw.in/journal/2006/27-28/
commentary/restructuring-watershed-development-programmes.html.
  1756. Kerr, J., 2007: Watershed management: Lessons from common property theory. Int. J. Commons, 1, 89–109, doi:10.18352/ijc.8.
  1757. Bhalla, R.S., N.W. Pelkey, and K.V. Devi Prasad, 2011: Application of GIS for evaluation and design of watershed guidelines. Water Resour. Manag., 25, 113–140, doi:10.1007/s11269-010-9690-0.
  1758. Chandrasekhar, C.P., J. Ghosh, and A. Roychowdhury, 2006: The ‘demographic dividend’and young India’s economic future. Econ. Polit. Wkly., 5055–5064.
  1759. Deshpande, R.S., 2008: Watersheds: Putting the cart before the horse. Econ. Polit. Wkly., 43, 74–76.
  1760. Bharucha, Z.P., D. Smith, and J. Pretty, 2014: All paths lead to rain: Explaining why watershed development in India does not alleviate the experience of water scarcity. J. Dev. Stud., 50, 1209–1225, doi:10.1080/00220388.2014.
928699.
  1761. Calder, I., A. Gosain, M.S.R.M. Rao, C. Batchelor, M. Snehalatha, and E. Bishop, 2008: Watershed development in India. 1. Biophysical and societal impacts. Environ. Dev. Sustain., 10, 537–557, doi:10.1007/s10668-006-9079-7.
  1762. O’Brien, K. et al., 2004: Mapping vulnerability to multiple stressors: Climate change and globalization in India. Glob. Environ. Chang., 14, 303–313, doi:10.1016/J.GLOENVCHA.2004.01.001.
  1763. Samuel, A. et al., 2007: Watershed Development in Maharashtra: Present Scenario and Issues for Restructuring the Programme. http://www.indiawater
portal.org/articles/watershed-development-maharashtra-present-scenario-
and-issues-restructuring-programme. Society for Promoting Participative 
Ecosystem Management (SOPPECOM), Pune, India, 158 pp.
  1764. Samuel, A., and K.J. Joy, 2018: Changing land use, agrarian context and rural transformation. In: India’s Water Futures [Joy, K. J and N. Janakarajan (eds.)]. Routledge, India, pp. 57–78.
  1765. Mosase, E., and L. Ahiablame, 2018: Rainfall and temperature in the Limpopo River Basin, southern Africa: Means, variations, and trends from 1979 to 2013. Water, 10, 364, doi:10.3390/w10040364.
  1766. Jury, M.R., 2016: Climate influences on upper Limpopo River flow. Water SA, 42, 63, doi:10.4314/wsa.v42i1.08.
  1767. Christie, F., and J. Hanlon, 2001: Mozambique and the Great Flood Of 2000. Indiana University Press, Indiana, USA. ISBN 0-85255-858-9 (hardback); 0-85255-857-0 (paperback). 176 pp.
  1768. Engelbrecht, F.A. et al., 2011: Multi-scale climate modelling over southern Africa using a variable-resolution global model. Water Research Commission 40-year Celebration Conference, Kempton Park, 37, 647–658. http://dx.doi.org/10.4314/wsa.v37i5.2.
  1769. Spaliviero, M., M. De Dapper, and S. Maló, 2014: Flood analysis of the Limpopo River Basin through past evolution reconstruction and a geomorphological approach. Nat. Hazards Earth Syst. Sci., 14, 2027–2039, doi:10.5194/nhess-14-2027-2014.
  1770. Kahinda, J.M., Meissner, R., and Engelbrecht, F.A., 2016: Implementing integrated catchment management in the upper Limpopo River Basin: 
A situational assessment. Phys. Chem. Earth, 93, 104–118, doi:10.1016/j.pce.2015.10.003.
  1771. Alba, R., A. Bolding, and R. Ducrot, 2016: The politics of water payments and stakeholder participation in the Limpopo River Basin, Mozambique Raphaëlle Ducrot. Water Altern., 9, 569–587.
  1772. Albanito, F. et al., 2016: Carbon implications of converting cropland to bioenergy crops or forest for climate mitigation: A global assessment. GCB Bioenergy, 8, 81–95, doi:10.1111/gcbb.12242.
  1773. Gbetibouo, G.A., R.M. Hassan, and C. Ringler, 2010: Modelling farmers’ adaptation strategies for climate change and variability: The case of the Limpopo Basin, South Africa. Agrekon, 49, 217–234, doi:10.1080/03031853.2010.491294.
  1774. Machethe, C.L. et al., 2004: Smallholder Irrigation and Agricultural Development in the Olifants River Basin of Limpopo Province: Management Transfer, Productivity, Profitability and Food Security Issues. Report to the Water Research Commission on the Project ‘Sustainable Local Management of Smallholder Irrigation’, University of the North, Manitoba, Canada, 112 pp.
  1775. Spaliviero, M., M. De Dapper, C.M. Mannaerts, and A. Yachan, 2011: Participatory approach for integrated basin planning with focus on disaster risk reduction: The case of the Limpopo river. Water (Switzerland), 3, 737–763, doi:10.3390/w3030737.
  1776. van der Zaag, P., and H. Savenije, 1999: The management of international waters in EU and SADC compared. Phys. Chem. Earth, Part B Hydrol. Ocean. Atmos., 24, 579–589, doi:10.1016/S1464-1909(99)00048-9.
  1777. Kahinda, J.M., Meissner, R., and Engelbrecht, F.A., 2016: Implementing integrated catchment management in the upper Limpopo River Basin: 
A situational assessment. Phys. Chem. Earth, 93, 104–118, doi:10.1016/j.pce.2015.10.003.
  1778. Spaliviero, M., M. De Dapper, C.M. Mannaerts, and A. Yachan, 2011: Participatory approach for integrated basin planning with focus on disaster risk reduction: The case of the Limpopo river. Water (Switzerland), 3, 737–763, doi:10.3390/w3030737.
  1779. D’Odorico, P. et al., 2013: Global desertification: Drivers and feedbacks. Adv. Water Resour., 51, 326–344, doi:10.1016/j.advwatres.2012.01.013.
  1780. Zhang, Z., and D. Huisingh, 2018: Combating desertification in China: Monitoring, control, management and revegetation. J. Clean. Prod., 182, 765–775, doi:10.1016/J.JCLEPRO.2018.01.233.
  1781. Zucca, C., R. Della Peruta, R. Salvia, S. Sommer, and M. Cherlet, 2012: Towards a World Desertification Atlas. Relating and selecting indicators and data sets to represent complex issues. Ecol. Indic., 15, 157–170, doi:10.1016/J.ECOLIND.2011.09.012.
  1782. Kaspari, S., T.H. Painter, M. Gysel, S.M. Skiles, and M. Schwikowski, 2014: Seasonal and elevational variations of black carbon and dust in snow and ice in the Solu-Khumbu, Nepal and estimated radiative forcings. Atmos. Chem. Phys., 14, 8089–8103, doi:10.5194/acp-14-8089-2014.
  1783. Qian, Y. et al., 2015: Light-absorbing particles in snow and ice: Measurement and modeling of climatic and hydrological impact. Adv. Atmos. Sci., 32, 64–91, doi:10.1007/s00376-014-0010-0.
  1784. Painter, T.H., M.G. Flanner, G. Kaser, B. Marzeion, R.A. VanCuren, and 
W. Abdalati, 2013: End of the Little Ice Age in the Alps forced by industrial black carbon. Proc. Natl. Acad. Sci., 110, 15216–15221, doi:10.1073/PNAS.1302570110.
  1785. Olsson, L., M. Opondo, P. Tschakert, A. Agrawal, S.H. Eriksen, S. Ma, L.N. Perch, and S.A. Zakieldeen, 2014: Livelihoods and Poverty. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White, (eds.)]. Cambridge University Press, 
New York, USA. pp. 793–832.
  1786. FAO, 2016: Trees, Forests and Land Use in Drylands: The First Global Assessment Preliminary Findings. Food and Agriculture Organization of the United Nations, Rome, Italy, 31 pp.
  1787. Abatzoglou, J.T., S.Z. Dobrowski, S.A. Parks, and K.C. Hegewisch, 2018: TerraClimate, a high-resolution global dataset of monthly climate and climatic water balance from 1958–2015. Sci. Data, 5, 170191, doi:10.1038/sdata.2017.191.
  1788. Abatzoglou, J.T., S.Z. Dobrowski, S.A. Parks, and K.C. Hegewisch, 2018: TerraClimate, a high-resolution global dataset of monthly climate and climatic water balance from 1958–2015. Sci. Data, 5, 170191, doi:10.1038/sdata.2017.191.
  1789. FAO, 2016: Trees, Forests and Land Use in Drylands: The First Global Assessment Preliminary Findings. Food and Agriculture Organization of the United Nations, Rome, Italy, 31 pp.
  1790. van der Esch, S. et al., 2017: Exploring Future Changes in Land Use and Land Condition and the Impacts on Food, Water, Climate Change and Biodiversity: Scenarios for the UNCCD Global Land Outlook. Policy Report, PBL Netherlands Environmental Assessment Agency, the Hague, Netherlands, 115 pp.
  1791. Rahman, M.S., 2013: Climate change, disaster and gender vulnerability: A study on two divisions of Bangladesh. Am. J. Hum. Ecol., 2, 72–82, doi:10.11634/216796221504315.
  1792. Le, Q., C. Biradar, R. Thomas, C. Zucca, and E. Bonaiuti, 2016a: Socio-ecological context typology to support targeting and upscaling of sustainable land management practices in diverse global Dryland. In: Proceedings of the International Congress on Environmental Modeling and Software, Paper 45, July 2016, Toulouse, France [Sauvage, S., J. Sanchez-Perez and A. Rizzoli (eds.)]. Brigham Young University, Utah, USA.
  1793. Franks, P.J. et al., 2013: Sensitivity of plants to changing atmospheric CO2 concentration: From the geological past to the next century. New Phytol., 3 197, 1077–1094, doi:10.1111/nph.12104.
  1794. Burrell, A.L., J.P. Evans, and Y. Liu, 2017: Detecting dryland degradation using Time Series Segmentation and Residual Trend analysis (TSS-RESTREND). Remote Sens. Environ., 197, doi:10.1016/j.rse.2017.05.018.
  1795. Burrell, A.L., J.P. Evans, and Y. Liu, 2018: The impact of dataset selection on land degradation assessment. ISPRS J. Photogramm. Remote Sens., 146, 22–37, doi:10.1016/J.ISPRSJPRS.2018.08.017.
  1796. NIAB, 1997: Economic utilization of salt-affected soils In Twenty Five Years (1992–1997) of NIAB (Silver Jubilee Publication). Nuclear Institute for Agriculture and Biology, Faisalabad, Pakistan, pp. 123–151.
  1797. Akinshina, N., A. Azizov, T. Karasyova, and E. Klose, 2016: On the issue of halophytes as energy plants in saline environment. Biomass and Bioenergy, 91, 306–311, doi:10.1016/J.BIOMBIOE.2016.05.034.
  1798. Sanz, M.J. et al., 2017: Sustainable Land Management Contribution to Successful Land-Based Climate Change Adaptation and Mitigation. A Report of the Science-Policy Interface. United Nations Convention to Combat Desertification, Bonn, Germany, 170 pp.
  1799. Klein, R.J.T., G.F. Midgley, B.L. Preston, M. Alam, F.G.H. Berkhout, K. Dow, and M.R. Shaw, 2014: Adaptation Opportunities, Constraints, and Limits. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 899–943, doi:10.1017/CBO9781107415379.021.
  1800. Dow, K. et al., 2013: Limits to adaptation. Nat. Clim. Chang., 3, 305–307, doi:10.1038/nclimate1847.
  1801. Klein, R.J.T., G.F. Midgley, B.L. Preston, M. Alam, F.G.H. Berkhout, K. Dow, and M.R. Shaw, 2014: Adaptation Opportunities, Constraints, and Limits. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Field, C.B., V.R. Barros, D.J. Dokken, K.J. Mach, M.D. Mastrandrea, T.E. Bilir, M. Chatterjee, K.L. Ebi, Y.O. Estrada, R.C. Genova, B. Girma, E.S. Kissel, A.N. Levy, S. MacCracken, P.R. Mastrandrea, and L.L.White (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 899–943, doi:10.1017/CBO9781107415379.021.
  1802. Çarkaci, D.A., 1999: Examples of good practices of Turkey for combating wind erosion (examples of karapinar) the blooming desert Karapinar situation of desertification in Turkey, distribution of problem areas in Turkey, http://www.mgm.gov.tr/FTPDATA/arastirma/toz/Day3_99_acarkaci.pdf.
  1803. Türkeş, M., M. Turp, T. An, N. Ozturk, and M.L. Kurnaz, 2019: Impacts of climate change on precipitation climatology and variability in Turkey. In: Water Resources of Turkey [Harmancioglu, N.B. and D. Altinbilek, (eds.)]. Springer International Publishing, New York, USA. Pages 467–491.
  1804. Çarkaci, D.A., 1999: Examples of good practices of Turkey for combating wind erosion (examples of karapinar) the blooming desert Karapinar situation of desertification in Turkey, distribution of problem areas in Turkey, http://www.mgm.gov.tr/FTPDATA/arastirma/toz/Day3_99_acarkaci.pdf.
  1805. Türkeş, M., 2013: İklim Verileri Kullanılarak Türkiye’nin Çölleşme Haritası Dokümanı Hazırlanması Raporu. Orman ve Su İşleri Bakanlığı, Çölleşme ve Erozyonla Mücadele Genel Müdürlüğü Yayını, Ankara, Turkey, 57 pp.
  1806. Saifi, M., Boulghobra, N. and Fattoum, L., and M. Oesterheld, 2015: The Green Dam in Algeria as a tool to combat desertification. Planet@Risk, 3, 68–71, doi:10.1016/j.jaridenv.2006.01.021.
  1807. Berrahmouni, N., L. Laestadius, A. Martucci, D. Mollicone, C. Patriarca, and M. Sacande, 2016: Building Africa’s Great Green Wall: Restoring Degraded Drylands for Stronger and More Resilient Communities. Food and Agriculture Organization of the United Nations, Rome, Italy, 7 pp.
  1808. Bellard, C., W. Thuiller, B. Leroy, P. Genovesi, M. Bakkenes, and F. Courchamp, 2013: Will climate change promote future invasions? Glob. Chang. Biol., 19, 3740–3748, doi:10.1111/gcb.12344.
  1809. Bellard, C., W. Thuiller, B. Leroy, P. Genovesi, M. Bakkenes, and F. Courchamp, 2013: Will climate change promote future invasions? Glob. Chang. Biol., 19, 3740–3748, doi:10.1111/gcb.12344.
  1810. Rashid, M., S.H. Abbas, and A. Rehman, 2014: The Status of highly alien invasive plants in Pakistan and their impact on the ecosystem: A Review. Innovare J. Agric. Sci,. 2, 2–5.
  1811. Benmoussa, H., E. Luedeling, M. Ghrab, J. Ben Yahmed, and B. Mimoun, 2017: Performance of pistachio (Pistacia vera L.) in warming Mediterranean orchards. Environ. Exp. Bot., 140, 76–85, doi:10.1016/j. envexpbot.2017.05.007.
  1812. Tabieh, M. et al., 2015: Economic analysis of micro-catchment rainwater harvesting techniques in Jordan’s arid zones. Int. J. Appl. Environ. Sci., 10, 1205–1225.
  1813. Oweis, T.Y., 2017: Rainwater harvesting for restoring degraded dry agropastoral ecosystems: A conceptual review of opportunities and constraints in a changing climate. Environ. Rev., 25, 135–149, doi:10.1139/er-2016-0069.

Land Degradation