SAND AND DUST STORMS IN THE MIDDLE EAST AND NORTH AFRICA (MENA) REGION Sources, Costs, and Solutions Fall 2019 Environm nt, N tur l R sourc s & Blu Econom SAND AND DUST STORMS IN THE MIDDLE EAST AND NORTH AFRICA (MENA) REGION SOURCES, COSTS, AND SOLUTIONS © 2019 International Bank for Reconstruction and Development/The World Bank 1818 H Street NW Washington, DC 20433 Telephone: 202-473-1000 Internet: www.worldbank.org This work is a product of the staff of The World Bank with external contributions. The findings, interpretations, and conclusions expressed in this work do not necessarily reflect the views of The World Bank, its Board of Executive Direc- tors, or the governments they represent. The World Bank does not guarantee the accuracy of the data included in this work. The boundaries, colors, denomina- tions, and other information shown on any map in this work do not imply any judgment on the part of The World Bank concerning the legal status of any territory or the endorsement or acceptance of such boundaries. Rights and Permissions The material in this work is subject to copyright. The World Bank encourages dissemination of its knowledge, this work may be reproduced, in whole or in part, for noncommercial purposes as long as full attribution to this work is given. Attribution—Please cite the work as follows: World Bank. 2019. Sand and Dust Storms in the Middle East and North Africa Region—Sources, Costs, and Solutions. Washington, DC. Any queries on rights and licenses, including subsidiary rights, should be addressed to: World Bank Publications The World Bank Group 1818 H Street NW Washington, DC 20433 USA Fax: 202-522-2625 TABLE OF CONTENTS Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v Acronyms and Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi Executive Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii Trends and Sources of Dust Storms in MENA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Dust Hot Spots and Trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Sources and Drivers of Sand and Dust Storms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Impacts of Increased Dust Concentration and Dust Storms . . . . . . . . . . . . . . . . . . . . . . . . . . 10 Health Impacts of Dust Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 Environmental Impacts of Dust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Economic Costs of Dust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 Preventative Actions, Interventions, and Policies against Sand and Dust Storms . . . . . . . 19 Early Warning Systems for Dust Storm Prediction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 Technical Interventions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 National and Regional Government Policies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 Sources, Costs, and Solutions iii FIGURES Figure 1: Global Pattern of Dust Frequency Estimated from the Synoptic Present Weather Records for the Period of January 1974 to December 2012 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 Figure 2: Geographic Distribution of the Dust Atmospheric Load . . . . . . . . . . . . . . . . . . . . . . . . . 3 Figure 3: Time Series of Global Monthly Mean Dust Concentration and the Corresponding 95% Confidence Interval (in error bars) for the Period 1974–2012 . . . . . . . . . . . . . . . . . . . . . . . . . 3 Figure 4: Sand and Dust Storms Path and Source Clusters in MENA . . . . . . . . . . . . . . . . . . . . . . 5 Figure 5: Distribution of the Percentage Number of Days per Year with Dust Optical Depth > 0.2 over North Africa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 Figure 6: Distribution of the Percentage Number of Days per Year with Dust Optical Depth > 0.2 over the Middle East . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 Figure 7: Global PM10 Levels and DALYs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 Figure 8: Annual PM10 Concentration, Deaths, and DALYs in MENA Countries . . . . . . . . . . . . . 12 Figure 9: Technologies for Sand and Dust Storm Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 TABLES Table 1: Key Physical Factors Influencing Wind Erosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 Table 2: Short-term and Long-term Impacts of Sand and Dust Storms . . . . . . . . . . . . . . . . . . 16 Table 3: Welfare Losses from Ambient PM2.5 by Region (2011 US$ billions— PPP adjusted) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Table 4: Mean Annual PM2.5, and Total Deaths and Losses from Pollution by Country . . . . . . 18 Table 5: Institutions and Organizations with Dust Forecasting Programs . . . . . . . . . . . . . . . 22 BOXES Box 1: Indirect Human-Induced Factors That Contribute to Sand and Dust Storms . . . . . . . . 9 Box 2: Economic Cost Case Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Box 3: UN-Interagency Response to SDS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 Box 4: Regional Air Pollution Policies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 iv Sand and Dust Storms in the Middle East and North Africa (MENA) Region ACKNOWLEDGMENTS This report was prepared by Menaal Ebrahim under the guidance of Craig Meisner, Senior Environmental Economist of the World Bank. The team would like to thank their colleagues from the World Bank for their useful advice and support throughout this project: Benoit Blarel, Practice Manager, Lia Sieghart, Practice Manager, Tim Brown, Senior Natural Resources Management Specialist, Philippe Dardel, Senior Environmental Specialist, Raffaello Cervigni, Lead Environmental Economist, Paola Agostini, Lead Natural Resources Management Specialist, and Melissa Landesz, Senior Natural Resources Management Specialist. The team would also like to acknowledge the financial support of this study through the Program for Forests (PROFOR) Trust Fund. Sources, Costs, and Solutions v ACRONYMS AND ABBREVIATIONS ADB Asian Development Bank AEMET Meteorological State Agency of Spain AI Aerosol Index APINA Air Pollution Information Network for Africa ASEAN Association of Southeast Asian Nations AUD Australian Dollar CAWAS Centre for Atmosphere Watch and Services (China) CLRTAP Convention on Long-Range Transboundary Air Pollution CNY Chinese Yuan COPD Chronic Obstructive Pulmonary Disease DALYs Disability-Adjusted Life Years ECMWF European Center for Medium-range Weather Forecasting (UK) FAO Food & Agriculture Organization GDP Gross Domestic Product IARS Inertial Altitude Reference System IASI Infrared Atmospheric Sounding Interferometer ICAO International Civil Aviation Organization IHME Institute of Health Metrics and Evaluation ITU International Telecommunication Union IUCN International Union for Conservation of Nature KMA Korea Meteorological Administration km2 Square kilometers MENA Middle East and North Africa MODIS Moderate Resolution Imaging Spectroradiometer NIES National Institute for Environmental Studies (Japan) NWS National Weather Service (USA) PM2.5 Particulate Matter (diameter less than 2.5 microns) PM10 Particulate Matter (diameter less than 10 microns) PEF Peak Expiratory Flow PPP Purchasing Power Parity RAPIDC Regional Air Pollution in Developing Countries RC Regional Centre SA Source Apportionment SACEP South Asia Cooperative Environment Programme SCS Soil Conservation Service (USA) SDS Sand and Dust Storm SDS-WAS Sand and Dust Storm Warning Advisory and Assessment System SMS Short Message Service UAE United Arab Emirates UAV Unmanned Aerial Vehicle UN United Nations UNCCD United Nations Convention to Combat Desertification UNDP United Nations Development Programme UNECE United Nations Economic Commission for Europe UNEP United Nations Environment Programme vi Sand and Dust Storms in the Middle East and North Africa (MENA) Region UNESCAP United Nations Economic and Social Commission for Asia and Pacific UNESCWA United Nations Economic and Social Commission for Western Asia UN-Habitat United Nations Human Settlement Programme UNITAR United Nations Institute for Training and Research USD United States Dollars WHO World Health Organization WMO World Meteorological Organization WRF Weather Research and Forecasting WSN Wireless Sensor Network Sources, Costs, and Solutions vii EXECUTIVE SUMMARY Dust storms are capable of transporting sediment over thousands of kilome- ters, but due to the Middle East and North Africa (MENA) region’s proximity to the Sahara Desert, the region is one of the dustiest in the world. Dust storms are transboundary, which has important implications for their mitigation, as effects are felt in different countries and even regions than their source of origin. North African dust is trans- ported to as far away as the Amazon Forest, North America, Europe, and China. The Sahara Desert is undoubtedly the biggest dust source, as its dust emissions are about four times as much as Arabian deserts. North Africa, the Middle East, South West Asia, and North East Asia are the regions with the highest dust frequencies and highest Aerosol Index (AI) values. The highest density of dust sources in the Middle East is found in northern Iraq between the Tigris and Euphrates rivers and along the Syria-Iraq border. Dust sources in the region are also generally found in areas with extensive desert cover, low population densities, and sparse agriculture concentrated along river valleys. In terms of occurrences of dust storms, Sudan, Iraq, Saudi Arabia, and the Persian Gulf report the greatest number of dust storms overall. While natural sources such as the Sahara are the main contributors to dust storms in MENA, land-use changes and human-induced climate change has added anthropogenic sources as well. There are about three times as many natural dust sources as anthropogenic dust sources, however due to land use changes in the past few decades, anthropogenic sources have increased. Most North African dust storms originate from natural sources such as the Sahara, but there are some anthropogenic sources too. For instance, Southern Sahel, the Atlas Mountains, and the Mediterranean coast sources are overwhelmingly anthropogenic. The Middle East region also experiences dust storms from a mix of natural and anthropogenic sources. The Aral Sea is an active dust source, as well as dry riverbeds in Saudi Arabia. There is a cluster of anthropogenic and hydrologic sources along the Jordan River, par- ticularly on the east side. In addition, Iran has prominent dust sources such as large salty lakes and deserts, but the northwestern part is anthropogenic. Although much dust in the Middle East derives from local sources, substantial amounts of dust come from the Sahara. Like sources, drivers of sand and dust storms are also natural and anthropo- genic, as both wind speed and land management can cause them. Wind erosion is the main driver of sand storms and dust emissions in all systems. Essentially, erodibility of sur- face material coupled with aridity that limits protective effects of vegetation is a natural driving factor of sand and dust storms. For instance, the major dust storm event in 2015 in the Middle East has been attributed to the wind and arid conditions in the area rather than man-made fac- tors. However, there are many man-made drivers of wind erosion too. Human-induced land degradation is a driver of wind erosion and a major contributor toward sand and dust storms, as it exposes degraded and dry surfaces with a long wind fetch. Besides resource use and man- agement in drylands, practices that disrupt the hydrology and protection provided by land in general also contribute to sand and dust storms. In addition, land management practices that result in deforestation and clearance of vegetation will lead to an increase of wind velocity, as well as reduce the entrapment of particles. The complementary piece titled ‘Sustainable Land Management and Restoration in the Middle East and North Africa Region—Issues, Challenges, and Rec- ommendations’ provides details on land degradation severity, drivers, and impacts and provides insights into the human-induced drivers of dust storms in MENA. viii Sand and Dust Storms in the Middle East and North Africa (MENA) Region Dust deposition has wide-ranging health impacts, such as causing and aggra- vating asthma, bronchitis, respiratory diseases, and infections and lung cancer. Populations far from the source regions are exposed to a wide range of air quality– related health problems when long-range atmospheric transport carries dust. For instance, African dust transported to the Caribbean and Florida has deteriorated air quality standards in those areas and makes up half of South Florida’s airborne particles in the summer. Poor air quality and dust cause numerous health problems, both near the dust storm and thou- sands of kilometers away. Inhalation of fine particles can cause or aggravate diseases such as asthma, bronchitis, emphysema, and silicosis. Chronic exposure can be linked to respiratory disease, lung cancer, and acute lower respiratory infections. Apart from devastating health impacts, dust also impacts the environment, agriculture, transport, and infrastructure. For the environment, dust can have both negative and positive effects. Dust storms have some positive global impacts due to their trans- boundary nature and the importance of dust in global climate and terrestrial and biogeo- chemical cycling. For instance, dust fertilizes and sustains both oceans and forests, playing a huge part in the earth’s biogeochemical cycles. While dust boosts primary productivity of oceans, it could have damaging effects on coral reefs. In addition, dust has been also associ- ated with leading to and exacerbating climatic events such as storms, droughts, and the melt- ing of glaciers. Dust deposition and dust storms are also associated with many other costs such as crop damage, livestock mortality, infrastructure damage, and interruption of transport. Globally, welfare losses from dust are approximately 3.6 trillion USD, where costs are about 150 billion USD and over 2.5 percent of Gross Domestic Prod- uct (GDP) on average in MENA. Dust storm costs range from negative health impacts to reducing crop yields to lowering property values to steering talented workers away from polluted places. The World Health Organization estimates that 7 million people die from poor air quality every year, which is at least partly attributed to dust. A study prepared by the World Bank and the Institute of Health Metrics and Evaluation (IHME) calculated welfare losses from ambient PM2.5 pollution for each country, where welfare losses represent the cost of premature mortality. Global welfare losses from premature mortality are large and increasing from 2.2 trillion in 1990 to 3.6 trillion USD in 2013. For MENA, dust concentra- tion and storms cost MENA over 150 billion USD annually and over 2.5 percent of GDP for most countries in the region. According to the UN, about 13 billion USD are lost every year from dust storms alone in the MENA region and welfare losses from PM2.5 alone were about 141 billion USD in 2013. The biggest welfare losses were incurred by Egypt, Iran, and Pakistan. Besides investing in early warning systems, governments all over the world are designing policies to mitigate the impact of sand and dust storms, both at national and regional levels. Devastating impacts from sand and dust storms in the Americas, MENA region, and East Asia have encouraged governments to enforce many large-scale initiatives and plans. In many instances, these initiatives also tackle land degra- dation, terrestrial biodiversity, and climate change mitigation. However, policies designed to mitigate the wider impacts of sand and dust storms, including many that are transboundary, are geographically patchy and have a much shorter history. Regional and international coop- eration among countries will lead to greater understanding of the transportation paths of dust storms, particle content, and their impacts. Eventually, regional action will also lead to reduced the occurrence of dust storms. Recent years have seen some regional air pollution policies emerge, but more collaboration is needed and should be sustained. Sources, Costs, and Solutions ix TRENDS AND SOURCES OF DUST STORMS IN MENA Sand and dust storms (SDS) are complex events with transboundary impacts. Sand and dust storms result from the erosion and transport of mineral sediments from land. Sand particles are larger than dust, but both are typically associated with dryland areas and can occur anywhere where there are dry unprotected sediments.1 They could lift large quan- tities of dust particles into the air and transport them hundreds or thousands of kilometers away.2 SDS occur because of interlinked direct and indirect drivers, divided into natural and anthropogenic sources. Concern on sand and dust storms is growing considering their huge impacts on the economy, human health, and the environment. DUST HOT SPOTS AND TRENDS The MENA region, which neighbors the Sahara Desert, is the dustiest region in the world. Nine regions contribute to the total global production of desert dust: North Africa (Sahara), South Africa, the Arabian Peninsula, Central Asia, Western China, Eastern China, North America, South America and Australia.3, 4 North Africa, the Middle East, South West Asia, and North East Asia are the regions with the highest dust frequencies, as observed from synoptic weather reports (Figure 1).5 Similarly, in terms of Aerosol Index (AI) hot spots, the Sahara and Asian deserts are dominant, whereas AI values are low in the Southern Hemisphere and the Americas.6 The dust observed in the Caribbean is trans- ported dust from the Sahara, while the dust observed in Mexico may be partly related to the dust activities in the Chihuahua Desert.7, 8 The Sahara Desert is undoubtedly the biggest dust source, as its dust emissions are about four times as much as Arabian deserts. Dust can travel thousands of kilometers, as North African dust is transported as far as the Caribbean. Dust storms are capable of transporting sediment over thou- sands of kilometers. Dust storms are transboundary, which has important implications for their 1Middleton and Goudie, Desert Dust in the Global System. 2Zoljoodi, Didevarasl, and Saadatabadi, “Dust Events in the Western Parts of Iran and the Relationship with Drought Expansion over the Dust-Source Areas in Iraq and Syria.” 3Prospero et al., “Environmental Characterization of Global Sources of Atmospheric Soil Dust Identified with the Nimbus 7 Total Ozone Mapping Spectrometer (TOMS) Absorbing Aerosol Product.” 4Tanaka and Chiba, “A Numerical Study of the Contributions of Dust Source Regions to the Global Dust Budget.” 5Shao, Klose, and Wyrwoll, “Recent Global Dust Trend and Connections to Climate Forcing.” 6Middleton and Goudie, Desert Dust in the Global System. 7Prospero, “Long-Range Transport of Mineral Dust in the Global Atmosphere.” 8Liu et al., “CALIPSO Lidar Observations of the Optical Properties of Saharan Dust.” Sources, Costs, and Solutions 1 FIGURE 1: GLOBAL PATTERN OF DUST FREQUENCY ESTIMATED FROM THE SYNOPTIC PRESENT WEATHER RECORDS FOR THE PERIOD OF JANUARY 1974 TO DECEMBER 2012 Source: Shao et al., 2013. mitigation, as effects are felt in different countries and even regions than their source of origin. For instance, dust from China can reach the European Alps, after being transported across the Pacific and Atlantic Oceans over 13 days.9 Dust from Central Asia and China reaches Korea, Japan, the Pacific Islands, and North America.10 Based on estimates made by Tanaka and Chiba (2006), Figure 2 shows the geographical distribution of the desert dust atmospheric loads, in mg m−2.11 North African dust is transported to as far away as the Amazon Forest, North America, Europe, and China. The westward dust movement from the Sahara is the largest flow, accounting for 30–50 percent of the output. For example, transport to the Carib- bean, where 20 million tons of Saharan dust are deposited annually, typically takes 5 to 7 days. While dust emissions have generally been high and increased over the last cen- tury, the past two decades have not seen a rise in emissions from the North Africa region. Simulations suggest that global annual dust emissions have increased by 25 to 50 percent over the last century due to a combination of land use and climate changes. Sand and dust storm frequency and severity have increased in recent decades in some areas but decreased in other areas. However, a recent analysis by Shao et al. (2013) revealed that over the period 1984–2012, the global mean of near-surface dust concentration decreased at 1.2 percent per year (Figure 3).12 This decrease is mainly due to reduced dust activities in North Africa, accompanied by reduced activities in Northeast Asia, South America, and South Africa. This could be attributed to recovery of vegetation because of rainfall following the droughts in the 1980s, leading to a reduction in wind.13 However other studies conclude that a reduction in wind cannot be directly linked to changes in land use.14 9Grousset et al., “Case Study of a Chinese Dust Plume Reaching the French Alps.” 10Middleton and Goudie, Desert Dust in the Global System. 11Tanaka and Chiba, “A Numerical Study of the Contributions of Dust Source Regions to the Global Dust Budget.” 12Shao, Klose, and Wyrwoll, “Recent Global Dust Trend and Connections to Climate Forcing.” 13Cowie, Knippertz, and Marsham, “Are Vegetation-Related Roughness Changes the Cause of the Recent Decrease in Dust Emission from the Sahel?” 14Ridley, Heald, and Prospero, “What Controls the Recent Changes in African Mineral Dust Aerosol across the Atlantic?” 2 Sand and Dust Storms in the Middle East and North Africa (MENA) Region FIGURE 2: GEOGRAPHIC DISTRIBUTION OF THE DUST ATMOSPHERIC LOAD Source: De Longueville et al., 2010.15 FIGURE 3: TIME SERIES OF GLOBAL MONTHLY MEAN DUST CONCENTRATION AND THE CORRESPONDING 95% CONFIDENCE INTERVAL (IN ERROR BARS) FOR THE PERIOD 1974–2012 Source: Shao et al., 2013. 15De Longueville et al., “What Do We Know about Effects of Desert Dust on Air Quality and Human Health in West Africa Compared to Other Regions?” Sources, Costs, and Solutions 3 Countries in the Middle East experience varying frequencies of dust storms depending on the time of year. The Middle East region is a notable dust hot spot, espe- cially during the summer months when the dust storms in the region are often associated with Shamal winds.16 Sudan, Iraq, Saudi Arabia, and the Persian Gulf report the greatest number of dust storms overall.17 In the summer months, Iran, Iraq, Syria, the Persian Gulf, and the southern Arabian Peninsula experience the most dust storms. In western Iraq and Syria, Jordan, Lebanon, northern Israel, northern Arabian Peninsula, and southern Egypt they occur mainly in the spring, while in southern Israel and in the Mediterranean parts of northern Egypt, they occur in winter and spring.18 Sand and dust storms in MENA are determined by numerous climate systems and pathways. There are a variety of climate systems that govern the distribution of sand and dust storm events in the MENA region such as the Siberian, polar, and monsoon cyclones, and the depressions in the non-summer months. In MENA, most of dust storm systems can be classified into Summer Shamal and frontal dust storms.19 Shamal dust storms usually occur across Iraq, Kuwait, western part of Khuzestan plain, and some parts of Ara- bian Peninsula, whereas frontal dust storms occur across Jordan, Israel, and the northern Arabian Peninsula.20 There are six main sand and dust storm paths dominated by the cli- mate in MENA (Figure 4). The first path originates from the Mediterranean Sea passing over Cyprus and enters Syria. The second path is under the control of a high-pressure system over east of Europe.21 The third path comes from south of the Mediterranean Sea or coastal of northern Africa and always strikes south of Syria or the north border of Jordan and Saudi Arabia. The fourth path is from north of Africa which usually passes across Egypt, north of the Red Sea, and blows toward southeast in Saudi Arabia.22 The fifth path is also located in the depressions in north of Africa. The last path originates from Sistan Plain at the Iran– Afghanistan border which is controlled by anticyclone over central Asia. Air masses from the Mediterranean Sea are important factors for the generation of sand and dust storms which cover about 70 percent dust storm events.23 SOURCES AND DRIVERS OF SAND AND DUST STORMS SOURCES OF SAND AND DUST STORMS Sand and dust storm sources and drivers are both natural and anthropogenic. There is need to distinguish drivers of sand and dust storms from natural sources, which sup- ply most of the global dust emissions and anthropogenic sources. However natural ecosys- tems are increasingly being subject to human pressure, which may intensify their importance as source areas in the future.24 Although there is currently much uncertainty on the mag- nitude of human activity on sand and dust storms, disturbance of natural systems through human pressure is highly likely to increase in the coming decades, including through human-­ induced climate change. 16Choobari, Zawar-Reza, and Sturman, “The Global Distribution of Mineral Dust and Its Impacts on the Climate System.” 17Furman, “Dust Storms in the Middle East.” 18Ibid. 19Hamidi, Kavianpour, and Shao, “Synoptic Analysis of Dust Storms in the Middle East.” 20Middleton, “Dust Storms in the Middle East.” 21Hamidi, Kavianpour, and Shao, “Synoptic Analysis of Dust Storms in the Middle East.” 22Wilderson, “Dust and Sand Forecasting in Iraq and Adjoining Countries.” 23Cao et al., “Identification of Dust Storm Source Areas in West Asia Using Multiple Environmental Datasets.” 24Assessment, Millennium Ecosystem. 4 Sand and Dust Storms in the Middle East and North Africa (MENA) Region FIGURE 4: SAND AND DUST STORMS PATH AND SOURCE CLUSTERS IN MENA Source: Cao et al., 2015. Globally, there are about three times as many natural dust sources as anthro- pogenic dust sources; however, due to land use changes in the past few decades, anthropogenic sources have increased. There are three dust source types: hydrologic, dust linked to various water features as discussed above; natural, dust emitted from land surfaces where land use is less than 30 percent; and anthropogenic, sources where land use exceeds 30 percent. North Africa accounts for 55 percent of global dust emissions with only 8 percent being anthropogenic, mostly from the Sahel. Hydrologic dust sources (e.g., ephem- eral water bodies) account for 31 percent worldwide; 15 percent of them are natural while 85 percent are anthropogenic. Overall, natural dust sources globally account for 75 percent of emissions and anthropogenic sources account for the rest. Most North African dust comes from natural sources such as the Sahara, with some anthropogenic sources. Southern Sahel sources are overwhelmingly anthropogenic (locations 1 to 5), whereas the Sahara is the most significant natural Sources, Costs, and Solutions 5 FIGURE 5: DISTRIBUTION OF THE PERCENTAGE NUMBER OF DAYS PER YEAR WITH DUST OPTICAL DEPTH > 0.2 OVER NORTH AFRICA Source: Ginoux et al., 2012.25 source (locations 6 to 11) (Figure 5). This could be explained by the fact that agricultural and grazing activities in regions with some rainfall are confined to relatively localized areas around point sources of water, and most agricultural and grazing activity takes place in wet- ter areas.26 Analysis of thousands of years of dust deposition in the mouth of the Senegal River showed a sharp increase in deposition after the advent of commercial agriculture in the Sahel, about 200 years ago.27 The sources in the Atlas Mountains (locations 20 to 23) and along the Mediterranean coast (e.g., location 19) are also mostly anthropogenic. Outside the Sahel, the major sources are natural.28 These include major depressions, large basins with sand seas, ephemeral lakes, and the Nile River Basin. 25Ginoux et al., “Global-Scale Attribution of Anthropogenic and Natural Dust Sources and Their Emission Rates Based on MODIS Deep Blue Aerosol Products.” 26Prospero et al., “Environmental Characterization of Global Sources of Atmospheric Soil Dust Identified with the Nimbus 7 Total Ozone Mapping Spectrometer (TOMS) Absorbing Aerosol Product.” 27Mulitza et al., “Increase in African Dust Flux at the Onset of Commercial Agriculture in the Sahel Region.” 28Prospero et al., “Environmental Characterization of Global Sources of Atmospheric Soil Dust Identified with the Nimbus 7 Total Ozone Mapping Spectrometer (TOMS) Absorbing Aerosol Product.” 6 Sand and Dust Storms in the Middle East and North Africa (MENA) Region FIGURE 6: DISTRIBUTION OF THE PERCENTAGE NUMBER OF DAYS PER YEAR WITH DUST OPTICAL DEPTH > 0.2 OVER THE MIDDLE EAST Source: Ginoux et al., 2012. The Middle East shows a complex mixture of natural and anthropogenic sources. The Aral Sea was formerly one of the largest lakes in the world (area 68,000 km2) but is now reduced to 10 percent of its original size. Large areas of the Aral Sea are now active dust sources, in agreement with in situ measurements.29 There is an extensive area of anthropogenic sources, mixed with hydrologic sources, in Saudi Arabia (location 9) essen- tially aggregated around the dry riverbeds (Figure 6). The dust from the region between the Tigris and Euphrates is mapped as natural in Iraq and anthropogenic in Syria. There is a cluster of anthropogenic and hydrologic sources along the Jordan River, particularly on the east side (location 10). Iran has prominent dust sources such as large salty lakes and deserts, but the northwestern part is anthropogenic. Although much dust in the Middle East derives from local sources, substantial amounts of dust come from the Sahara. 29Wiggs et al., “The Dynamics and Characteristics of Aeolian Dust in Dryland Central Asia.” Sources, Costs, and Solutions 7 DRIVERS OF SAND AND DUST STORMS Wind erosion is the main natural driver of dust storms, which is also depen- dent on other climate and land characteristics. Wind is the main driver of sand storms and dust emissions in all systems. Specific synoptic meteorological conditions that produce winds vary in different regions.30 Globally about 32 million km2 of land is suscepti- ble to wind erosion, with 17 million km2 having high or very high susceptibility.31 Erodibility of surface material coupled with aridity that limits protective effects of vegetation is what usually defines natural dust sources and dust storms. For instance, the major dust storm event in 2015 in the Middle East has been attributed to the wind and arid conditions in the area rather than man-made factors.32 Besides wind speed, other land form characteristics also determine wind erosivity (Table 1). Considering these factors, the biggest dust sources are therefore usually inland drainage basins or depressions in arid areas, such as the Bodele Depression in the Sahara and the Taklamakan Desert in China.33, 34 TABLE 1: KEY PHYSICAL FACTORS INFLUENCING WIND EROSION Climate Sediment or Soil Vegetation Landform Wind speed (+) Soil type Type Surface roughness Wind direction Particle composition Coverage (–) Slope (–) Turbulence (+) Soil/sediment structure Density Ridge Precipitation (–) Organic matter (–) Distribution (+/–) Evaporation (+) Carbonates (–) Air temperature Bulk density Air pressure (+) Degree of aggregation (–) Freeze-thaw action Surface moisture (–) Source: UNEP, WMO, UNCCD, 2016; Shi et al., 2004; Middleton and Goudie, 2006.35 (+) indicates that the factor reinforces wind erosion, whereas (–) indicates that the factor has a protective effect, reducing wind erosion. (+/–) indicates the effect can be positive or negative depending on the processes involved. While climatic factors directly cause dust storms, there are several human-­ induced factors that can alter those climatic factors. Human-induced land degra- dation is a driver of wind erosion and a major contributor toward sand and dust storms, as it exposes degraded and dry surfaces with a long wind fetch. Besides resource use and man- agement in drylands, practices that disrupt the hydrology and protection provided by land in general also contribute to sand and dust storms. For instance, demand for water for urban areas or irrigation disturbs the hydrology of ephemeral lakes and playas. Building roads and other infrastructure that blocks the inflow of drainage waters is another contributor to the drying up of playas.36 Poor standards of crop management (e.g., related to soil fertility, seed quality, tillage, planting, and pest and disease control) that result in poor vegetation growth and soil cover increase risk of wind erosion. In addition, land management practices that result in deforestation and clearance of vegetation will lead to an increase of wind velocity, 30Knippertz and Stuut, “Mineral Dust.” 31Eswaran, Lal, and Reich, “Land Degradation.” 32Parolari et al., “Climate, Not Conflict, Explains Extreme Middle East Dust Storm.” 33Bullard et al., “Preferential Dust Sources.” 34Thomas, Arid Zone Geomorphology. 35UNEP, WMO, UNCCD., “Global Assessment of Sand and Dust Storms.” 36Gill, “Eolian Sediments Generated by Anthropogenic Disturbance of Playas.” 8 Sand and Dust Storms in the Middle East and North Africa (MENA) Region BOX 1: INDIRECT HUMAN-INDUCED FACTORS THAT CONTRIBUTE TO SAND AND DUST STORMS » Population increase and economic globalization leading to increased demands for food, feed, and other products » Failure of policy to recognize noneconomic ecosystem functions » Policies that unwittingly encourage unsustainable land management » Land use change to less sustainable uses » Use of prime agricultural land for urban development and waste disposal, thereby increas- ing pressure on marginal land » Subsistence farming » Lack of access to rural credit, extension services, and markets » Poverty » Insecure land tenure » Migration to fragile land » Climate change » War and insecurity Source: UNEP, WMO, UNCCD, 2016. as well as reduce the entrapment of particles. The complementary piece titled ‘Sustainable Land Management and Restoration in the Middle East and North Africa Region—Issues, Challenges, and Recommendations’ provides details on land degradation severity, drivers and impacts and offers insights into the man-made drivers of dust storms in MENA. There are also many other indirect drivers of sand and dust storms, such as population increase, weak land tenure, pov- erty, conflict, and climate change (Box 1). Sources, Costs, and Solutions 9 IMPACTS OF INCREASED DUST CONCENTRATION AND DUST STORMS Dust deposition has vast health, environmental, and economic impacts. Dust can contribute to numerous human health problems globally, especially in arid and semiarid regions. Inhalation of fine particles can cause or aggravate diseases such as asthma, bronchitis, emphysema, and silicosis. Chronic exposure can be linked to respiratory disease, lung cancer, and acute lower respiratory infections. For the environment, dust can have both negative and positive effects. Dust affects the climate system and can lead to intensifying drought conditions, but it also increases precipitation and provides nutrients to terrestrial ecosystems. In addition to health and environmental impacts, there are other short-term costs of dust such as crop damage, livestock mortality, infrastructure damage, and interruption of transport. Longer term costs include health problems, soil erosion, and disruption of global climate regulation. Mone- tizing these impacts can translate to hundreds of million dollars just from a single dust storm. HEALTH IMPACTS OF DUST DEPOSITION Dust storms often affect human life and health not only in the drylands but also in downwind regions. As discussed in the previous chapter, dust emitted from the North Africa region reaches as far as the rest of Africa, Middle East, Europe, Asia, the Carib- bean and the Americas, impacting air quality in those regions too. Dust from Asia is shown to contribute to aerosol loadings in western North America.37 African dust transported to the Caribbean and Florida has deteriorated air quality standards in those areas and makes up half of South Florida’s airborne particles in the summer.38 Therefore, populations far from the source regions are exposed to a wide range of air quality related health problems. Air quality, which is impacted by dust, is very poor in MENA. Airborne dust parti- cles, whether of natural origin and/or partially human from bush fires or practices that lead to desertification, affect human health through their impact on local and regional air quali- ties.39, 40 Airborne mineral dusts are respirable within size ranges of particles equal to or less 37Fairlie, Jacob, and Park, “The Impact of Transpacific Transport of Mineral Dust in the United States.” 38Prospero and Mayol-Bracero, “Understanding the Transport and Impact of African Dust on the Caribbean Basin.” 39Anuforom et al., “Inter-Annual Variability and Long-Term Trend of UV-Absorbing Aerosols during Harmattan Season in Sub-Saharan West Africa.” 40Sassen et al., “Saharan Dust Storms and Indirect Aerosol Effects on Clouds.” 10 Sand and Dust Storms in the Middle East and North Africa (MENA) Region than a diameter of 2.5 microns (PM2.5) and coarse particles equal to or less than 10 microns (PM10), as defined by the US Environmental Protection Agency. In the United States today, the following standards apply: the acceptable annual mean values of PM2.5 and PM10 are respectively 15 μg m−3 and 50 μg m−3 and the mean values over 24 h exceeding respectively 65 μg m−3 and 150 μg m−3 are considered to exceed the standards.41 Dust concentrations in the MENA region can reach well beyond these acceptable levels. Studies have shown that dust storms in the Middle East are characterized by high concentrations of particles with 2 to 20 μm diameter size, with more than 85 percent of them being less than 10 μm in diam- eter.42, 43 Increased dust concentrations can adversely affect health and even lead to death. Ambient fine particulate matter (PM2.5) exposure is currently considered the lead- ing environmental risk factor globally. Fine particulate matter can come from a variety of sources and its determination can be ascertained through a source apportionment (SA) study.44 Locations susceptible to high dust concentrations will have a larger contribution and in general will have a larger impact on premature mortality.45, 46 WHO’s environmen- tal burden of disease dataset that compiles PM10 levels and Disability Adjusted Life Years (DALYs) (which considers years lost to mortality) in all countries, shows a positive correlation between the two (Figure 7). It is estimated that PM2.5 exposure contributed to 4.1 million premature deaths in 2016.47 The situation is worse in areas that are prone to frequent dust events. Compared to other regions, MENA has one of the highest average PM2.5 and PM10 levels.48 Within MENA, populations in Iraq, Egypt, and Pakistan suffer disproportionately more in terms of premature deaths—as many as 30,000 deaths can be attributed to bad air quality (Figure 8). FIGURE 7: GLOBAL PM10 LEVELS AND DALYs 6.0 5.0 DALYs/1,000 capita per year 4.0 3.0 y = 0.011x + 0.2725 2.0 1.0 0 0 50 100 150 200 250 Annual PM10 (ug/m3) Source: Authors’ calculations based on WHO—Environmental burden of disease, 2004. 41Prospero, “Assessing the Impact of Advected African Dust on Air Quality and Health in the Eastern United States.” 42Perdue et al., “The Surgical Significance of Persian Gulf Sand.” 43Draxler et al., “Estimating PM Air Concentrations from Dust Storms in Iraq, Kuwait and Saudi Arabia.” 10 44The WHO maintains a database on source apportionment studies for particulate matter in the air (PM and PM ) at 10 2.5 https://www.who.int/quantifying_ehimpacts/global/source_apport/en/ 45Perez et al., “Coarse Particles from Saharan Dust and Daily Mortality.” 46Samoli et al., “Does the Presence of Desert Dust Modify the Effect of PM on Mortality in Athens, Greece?” 10 47Gakidou et al., “Global, Regional, and National Comparative Risk Assessment of 84 Behavioural, Environmental and Occupational, and Metabolic Risks or Clusters of Risks, 1990–2016.” 48World Health Organization, “Environmental Burden of Disease: Country Profiles.” Sources, Costs, and Solutions 11 FIGURE 8: ANNUAL PM10 CONCENTRATION, DEATHS, AND DALYs IN MENA COUNTRIES DALYsa *Annual **Urban /1000 Population PM10 Population Deaths Capital Country (’000) [µg/m3] (%) per Year per Year Afghanistan 24,076 27 16 400 0.3 Egypt 71,550 136 32 15,500 2 Iran 68,669 68 42 9,100 1.0 Iraq 27,456 167 58 10,300 6 Israel 6,574 53 80 1,400 1.1 Jordan 5,371 69 49 700 1.1 Kuwait 2,617 129 74 300 1.1 Lebanon 3,965 43 74 400 1.6 Libyan Arab Jamahiriya 5,799 121 85 1,800 3 Morocco 30,152 27 37 900 0.2 Oman 2,479 124 36 300 1.1 Pakistan 155,333 165 27 30,000 2.0 Qatar 764 57 65 <100 0.4 Saudi Arabia 23,047 91 40 2,500 1.1 Syrian Arab Republic 18,389 89 38 1,800 0.9 Tunisia 9,996 46 30 800 0.6 United Arab Emirates 3,947 109 70 200 0.7 Yemen 20,478 82 15 1,100 0.7 Source: WHO—Environmental burden of disease, 2004. *Urban-population-weighted average particulate matter less than 10 microns in diameters [mg/m3] (estimates or monitored, when available). **Percentage of urban population living in cities >100,000 and national capital. aFor Outdoor Air Pollution, DALYs consist only of years of life lost to premature mortality (YLL). RESPIRATORY/ASTHMA Respiratory illnesses are one of the main health impacts of dust. Airborne dust particles are transported via air inhaled through the nose or mouth and passed via the tra- chea to the lung tissues. Exposure to dust therefore contributes to respiratory disorders such as asthma, tracheitis, pneumonia, aspergillosis, allergic rhinitis, and nonindustrial silicosis.49 Dust has been strongly linked to Chronic Obstructive Pulmonary Disease (COPD), which is an umbrella term used to describe progressive lung diseases including emphysema, chronic bronchitis, and refractory (non-reversible) asthma. In Hong Kong, it was found that dust events had a significant adverse impact on emergency hospital admissions for COPD.50 Epi- demiological studies have also shown that increases in allergic rhinitis and daily admissions and clinical visits for allergic diseases such as asthma coincided with Asian dust storms.51, 52 Pneumonia admissions have also been significantly associated with Asian dust storms in Taipei.53 Desert dust also deteriorates pulmonary function. Recent studies have shown 49Derbyshire, “Natural Minerogenic Dust and Human Health.” 50Tam et al., “Effect of Dust Storm Events on Daily Emergency Admissions for Respiratory Diseases.” 51Chang et al., “Correlation of Asian Dust Storm Events with Daily Clinic Visits for Allergic Rhinitis in Taipei, Taiwan.” 52Kanatani et al., “Desert Dust Exposure Is Associated with Increased Risk of Asthma Hospitalization in Children.” 53Cheng et al., “Consequences of Exposure to Asian Dust Storm Events on Daily Pneumonia Hospital Admissions in Taipei, Taiwan.” 12 Sand and Dust Storms in the Middle East and North Africa (MENA) Region significantly reduced Peak Expiratory Flow (PEF) values and more increased PEF variability during dust days than during the control days in Korea.54, 55, 56 There is strong evidence on the adverse impacts of dust on asthma. Asthma is one of the world's leading noncommunicable diseases, and it affects about 334 million people each year.57 There is plenty of evidence that shows dust causes or exacerbates asth- matic conditions. Exposure to dust particles transported globally from desert storms is associ- ated worldwide with increased hospital admissions for childhood asthma and bronchitis, for example in Japan, Trinidad, and Texas.58, 59, 60 The highest prevalence of asthma has been reported in areas with desert dust storms events such as the MENA region.61, 62 In Greece, Saharan dust events have been associated with a 2.5 percent increase in pediatric asthma hospital admissions.63 Similarly, in Kuwait, dust storms led to an 8.4 percent increase in daily emergency asthma admissions over a period of five years, which was particularly evident among children.64 In Qatar, asthma cases are reported to increase by 30 percent during and shortly after very windy conditions.65 CARDIOVASCULAR Studies have found a positive correlation between dust events and cardiovas- cular illnesses, such as ischemic heart disease, cerebrovascular disease, and hypertension among others. Epidemiological studies have found positive relations between cardiovascular mortality and morbidity and dust storms. For instance, in Taiwan, a study of 39 Asian dust storm events found drastic increases in cardiopulmonary emergency visits when ambient PM10 concentrations were high and estimated that cardiovascular dis- eases, ischemic heart diseases, and cerebrovascular diseases during the Asian dust events increased by 26 percent, 35 percent, and 20 percent per event, respectively, compared to the pre-dust periods in Taiwan.66 Another study in China found a significant association between dust events and hypertension in men in Minqin China, and that the association of dust events and cardiovascular hospitalization was stronger in spring than in winter.67 There are some studies that have quantified cardiovascular health problems in MENA as well. The 2005 dust storm in Baghdad, Iraq, led to nearly 1,000 cases of suffocation.68 In Iran, dust storms caused a 1 percent increase in cardiovascular morbidity.69 54Gwack et al., “Effects of Asian Dust Events on Diurnal Variation of Peak Expiratory Flow Rate in Children with Bronchial Asthma and Healthy Children.” 55Yoo et al., “Acute Effects of Asian Dust Events on Respiratory Symptoms and Peak Expiratory Flow in Children with Mild Asthma.” 56Hong et al., “Asian Dust Storm and Pulmonary Function of School Children in Seoul.” 57Global Asthma Network, “Global Asthma Report.” 58Kanatani et al., “Desert Dust Exposure Is Associated with Increased Risk of Asthma Hospitalization in Children.” 59Gyan et al., “African Dust Clouds Are Associated with Increased Paediatric Asthma Accident and Emergency Admissions on the Caribbean Island of Trinidad.” 60Grineski et al., “Hospital Admissions for Asthma and Acute Bronchitis in El Paso, Texas.” 61Al Frayh et al., “Increased Prevalence of Asthma in Saudi Arabia.” 62Bener et al., “Genetic and Environmental Factors Associated with Asthma.” 63Samoli et al., “Acute Effects of Air Pollution on Pediatric Asthma Exacerbation: Evidence of Association and Effect Modification.” 64Thalib and Al-Taiar, “Dust Storms and the Risk of Asthma Admissions to Hospitals in Kuwait.” 65Teather et al., Examining the Links between Air Quality, Climate Change and Respiratory Health in Qatar. 66Chan et al., “Increasing Cardiopulmonary Emergency Visits by Long-Range Transported Asian Dust Storms in Taiwan.” 67Meng and Lu, “Dust Events as a Risk Factor for Daily Hospitalization for Respiratory and Cardiovascular Diseases in Minqin, China.” 68Middleton and Goudie, Desert Dust in the Global System. 69Delangizan and Jafari Motlagh, “Dust Phenomenon Affects on Cardiovascular and Respiratory Hospitalizations and Mortality: A Case Study in Kermanshah, during March-September 2010–2011.” Sources, Costs, and Solutions 13 OTHER INFECTIONS Infectious diseases, such as meningitis, conjunctivitis, and eye and skin infec- tions, are also known to be linked to increased dust concentrations. Meningo- coccal meningitis, also known as cerebrospinal meningitis, caused by the bacterium Neisseria meningitides, can cause large epidemics with fatality rates among cases.70 The largest epi- demic occurs in the African “meningitis belt,” a semiarid region spanning the Sahel from Senegal in the west to Ethiopia in the east, which has the highest rate of the disease.71 Dust storms from the Sahara and meningitis outbreaks are highly correlated and are perhaps linked through the Neisseria bacteria that need iron-laden dust to grow and become viral.72 Dust has also been linked to conjunctivitis, which is an inflammation of the conjunctiva and other ocular surfaces because of the reaction to an allergen.73 Exposure to desert dust could also lead to itchy eyes and skin rashes.74 Lastly, Asian dust is widely suspected to be an important factor in the pathogenesis of atopic dermatitis, which could be due to the fungi, mites, and other allergens contained in dust.75 ENVIRONMENTAL IMPACTS OF DUST Besides many of the negative impacts, dust fertilizes oceans and forests. Dust storms have some positive global impacts due to their transboundary nature and the impor- tance of dust in global climate and terrestrial and biogeochemical cycling.76 For instance, dust fertilizes and sustains both oceans and forests, playing a huge part in the earth’s biogeo- chemical cycles.77 Saharan dust fertilizes the Amazon forest, by replacing the phosphorous it loses from the basin. Similarly, Hawaiian rain forests receive nutrient inputs from dust from central Asia, which may sustain forest productivity over long time periods.78 While dust boosts primary productivity of oceans, it could have damaging effects on coral reefs. Dust provides nutrients to the surface and seabed of oceans, boost- ing primary productivity such as phytoplankton growth.79 Changes in dust fluxes to the ocean have the potential to modify ocean biogeochemistry.80 Research has suggested that dust deposition trends have increased ocean productivity by an estimated 6 percent in the past century.81 However, there is also a possibility that microorganisms, nutrients, trace met- als, and organic contaminants deposited in the dust on land and in oceans may play a role in the complex changes occurring on coral reefs worldwide.82 For instance, dust originating from Africa and Asia could therefore be adversely affecting coral reefs and other downwind ecosystems in the Americas. Dust has been also associated with leading to and exacerbating climatic events such as storms, droughts, and the melting of glaciers. Dust can affect climate by 70UNEP, WMO, UNCCD, “Global Assessment of Sand and Dust Storms.” 71World Health Organization, “Meningococcal Meningitis.” 72Noinaj, Buchanan, and Cornelissen, “The Transferrin–Iron Import System from Pathogenic N Eisseria Species.” 73Zhang et al., “A Systematic Review of Global Desert Dust and Associated Human Health Effects.” 74UNEP, WMO, UNCCD, “Global Assessment of Sand and Dust Storms.” 75Lee and Lee, “Effects of Asian Dust Events on Daily Asthma Patients in Seoul, Korea.” 76Ravi et al., “Aeolian Processes and the Biosphere.” 77Goudie, “Dust Storms.” 78Chadwick et al., “Changing Sources of Nutrients during Four Million Years of Ecosystem Development.” 79Jickells et al., “Air-Borne Dust Fluxes to a Deep Water Sediment Trap in the Sargasso Sea.” 80Aumont, Bopp, and Schulz, “What Does Temporal Variability in Aeolian Dust Deposition Contribute to Sea-Surface Iron and Chlorophyll Distributions?” 81Mahowald et al., “Observed 20th Century Desert Dust Variability.” 82Garrison et al., “African and Asian Dust.” 14 Sand and Dust Storms in the Middle East and North Africa (MENA) Region its effect on biogeochemical cycles, especially through effects on the ocean temperature and primary productivity and through indirect mechanisms from the dusts’ chemical reactivi- ty.83 Therefore, extreme events such as floods and droughts can be influenced by dust. For instance, dust has been linked to modifying tropical storms and cyclone intensities.84 Dust can also cause drought intensification, as dust loadings effect absorption and scattering of solar radiation and can alter the Earth’s radiative balance.85 Dust can affect precipitation indirectly too, through effects on convective activity due to altered temperature gradients. Glacial melt has also been linked to dust, as the deposition of mineral dust on glaciers has the potential to lower their surface albedo and speed up their melting.86 Dust deposition and storms are both a cause and symptom of land degra- dation. Wind erosion is one of the main land degradation processes, especially in dry- land regions.87 Wind erosion removes finer soil particles, which constitute the most active soil component in retaining nutrients and organic matter, resulting in soil degradation. The eroded material may damage crops and vegetation due to abrasion and sand burying young plants.88 Dust deposition has played a role in soil formation in many parts of the world, often at large distances from desert margins. The most striking example is the influence of aeolian processes on the formation of loess soils (unconsolidated silt), which occur extensively in North and South America, Central Asia, and China.89 Aeolian processes have also contrib- uted to forms of land degradation, such as soil salinization and alkalinity, through accumu- lation of soluble salt, and reduction of soil acidity through addition of carbonates.90 Thus dust entrainment (particle lifting by wind erosion) during dust events leads to long-term soil degradation, which is essentially irreversible. ECONOMIC COSTS OF DUST There are countless short-term and long-term impacts of dust pollution. Impacts range from negative health impacts to reducing crop yields to lowering property values to steering talented workers away from polluted places. As discussed in the previous sec- tion, sand dust affects crops and soil negatively. Sandblasting and burial of seedlings have an immediate negative effect on yields, and the loss of nutrient rich topsoil affects productivity in the long term.91 During dust storms, labor productivity and household incomes drop sharply, and millions of people are unable to reach work, and factories and offices close. Additionally, continued incidence of dust storms can also result in migration. Because of the Dust Bowl in the 1930s, millions of hectares of farmland became useless, and hundreds of thousands of people were forced to leave their homes.92 Other short-term impacts include livestock mortal- ity, infrastructure and transportation damage, and cost of clearing up sand (Table 2, Box 2). Global welfare losses from premature mortality are huge and increased from 2.2 trillion in 1990 to 3.6 trillion in 2013. Quantifying and monetizing dust impacts is difficult, as costs are wide-ranging and methods to calculate them are complex. Very few 83Singh et al., “Enhancement of Oceanic Parameters Associated with Dust Storms Using Satellite Data.” 84Evan et al., New Evidence for a Relationship between Atlantic Tropical Cyclone Activity and African Dust Outbreaks. 85Highwood and Ryder, “Radiative Effects of Dust.” 86Oerlemans, Giesen, and van den Broeke, “Retreating alpine glaciers: increased melt rates due to accumulation of dust (Vadret da Morteratsch, Switzerland).” 87Middleton and Goudie, Desert Dust in the Global System. 88Ravi et al., “Aeolian Processes and the Biosphere.” 89Muhs et al., “Identifying Sources of Aeolian Mineral Dust: Present and Past.” 90Middleton and Goudie, Desert Dust in the Global System. 91Behzad, Mineta, and Gojobori, “Global Ramifications of Dust and Sandstorm Microbiota.” 92Lee, Gill, and Mulligan, “The 1930s Dust Bowl.” Sources, Costs, and Solutions 15 TABLE 2: SHORT-TERM AND LONG-TERM IMPACTS OF SAND AND DUST STORMS Short-term Long-term Immediate human health problems (e.g., Cumulative human health problems (e.g., respiratory problems) and mortality bronchitis, cardiovascular disorders) Annual and perennial crop damage Soil erosion and reduced soil quality Livestock mortality Soil pollution through deposition of toxic biological materials (fungi, bacteria), heavy metals, or salts Infrastructural damage (e.g., buildings, Disruption of global climate regulation electricity and telephone structures, power (through feedbacks involving global facilities, solar farms, machinery, greenhouses) warming, ocean productivity and CO2 production, precipitation changes, global ice volume, sea level, hydrological cycle, and vegetation cover) Costs of clearing sand and dust from Migration infrastructure (e.g., roads, airports, dams, irrigation canals, flood control structures, ditches, power facilities) Interruption of transport (air, road, rail) and Decrease in household income communications; air and road traffic accidents Decline in labor productivity; office and business closure Source: Middleton and Goudie, 2006. studies have attempted to assess all costs associated with dust or a dust storm in a specific country (Box 2). However, just quantifying some of the most immense costs provides a mag- nitude of how big the dust problem is. The most significant is the cost of deaths and prema- ture mortality. The World Health Organization (WHO) estimates that 7 million people die from poor air quality every year, which is at least partly attributed to dust. A study prepared by the World Bank and the Institute of Health Metrics and Evaluation (IHME) calculated welfare losses from ambient PM2.5 pollution for each country, where welfare losses represent the cost of premature mortality.93 Globally, welfare losses increased from about 2.2 trillion to 3.6 trillion from 1990 to 2013 (Table 3). Interestingly, while Europe and North America have had the much higher welfare costs than other regions, they have not risen much over the years. On the other hand, welfare costs have at least doubled in all other regions. Dust concentration and storms cost MENA over 150 billion USD annually and over 2.5 percent of GDP for most countries in the region. The costs of dust pollu- tion and dust storms are significant in MENA. According to the UN, about 13 billion USD are lost every year due from dust storms alone in the MENA region.94 Additionally, welfare losses from PM2.5 were about 141 billion USD in 2013 in the MENA region, and an average of 2.5 percent of the GDP in MENA countries (Table 4).95 However, countries in MENA incur different costs depending on PM2.5 concentrations and the development level of the country. In absolute terms, the biggest welfare losses were incurred by Egypt, Iran, Pakistan, and Saudi Arabia. However, considering the economies, Egypt, Lebanon, Pakistan, and Yemen lost over 3 percent of their GDPs due to PM2.5 in 2013. 93World Bank and IHME, “The Cost of Air Pollution.” 94UNEP, “Sand and Dust Storms.” 95World Bank and IHME, “The Cost of Air Pollution.” 16 Sand and Dust Storms in the Middle East and North Africa (MENA) Region BOX 2: ECONOMIC COST CASE STUDIES The direct economic losses from the dust storm in May 1993 in China reached 550 million CNY.96 Driven by a cold air current from Siberia, a severe sandstorm occurred in northwest China in early May 1993. It moved southward from May 4 to 6, 1993, affecting a total area of 1.1 million square km. A total of 85 people died and 264 were injured, mostly pri- mary school children, 4,412 houses were destroyed, and 120,000 animals died or went missing. About 373,333 million hectares of crops were destroyed, over 2,000 km of irrigation ditches were buried, 16,300 ha of fruit trees were damaged, and thousands of greenhouses and plas- tic mulching sheds were broken; ground transportation (train and highways) was suspended, and telecommunications facilities were severely damaged in some areas. Many water resource back-up facilities, such as reservoirs, dams, catchments, underground canals, and flood control installations were filled up with sand silts. In 2002, Korea was estimated to have incurred costs of USD 4.6 billion, about 0.8 percent of its GDP from yellow dust.97 These costs included medical expenses, oppor- tunity costs, and industrial damage. The air transportation industry recorded sale losses due to flight cancellations caused by dust storms of USD 0.6 million. The total socioeconomic cost from yellow dust damage in South Korea in the year of 2002 is estimated at US$3,900 million at a minimum and US$7,300 million at a maximum, with an average of US$5,600 million, which is equivalent to 0.8% of GDP and US$117.00 per South Korean inhabitant. A large dust storm called Red Dawn that passed over the eastern coast of Austra- lia on 23 September 2009 is estimated to have cost AUD$299 million (with a range of AUD$293–A$313 million).98 Most of the costs were associated with household cleaning and associated activities. The study demonstrates some, but not all, of the major economic costs associated with wind erosion in Australia. Given the annual average cost of dust storms, the study suggested that AUD$9 million per year would be a conservative estimate of the level of investment required in rural areas for dust mitigation strategies, based on improved land management that could be justified to achieve a positive impact on soil conditions and reduce economic losses in rural towns and the more populous coastal cities. TABLE 3: WELFARE LOSSES FROM AMBIENT PM2.5 BY REGION (2011 US$ BILLIONS—PPP ADJUSTED) Region 1990 1995 2000 2005 2010 2013 East Asia and Pacific 273 366 458 668 1,065 1,387 Europe and Central Asia 1,247 1,172 1,129 1,232 1,188 1,170 Latin America and Caribbean 43 47 55 71 100 122 Middle East and North Africa 62 69 86 105 130 141 North America 483 503 527 518 451 431 South Asia 48 63 85 123 203 256 Sub-Saharan Africa 20 20 24 32 39 44 Total 2,176 2,240 2,364 2,749 3,176 3,551 Source: World Bank and IHME, 2016. 96Wang et al., “Analysis on the Formative Causes of Sand-Dust Storms in the Northwest China during 3–12 April 1994.” 97Jeong,“Socio-Economic Costs from Yellow Dust Damages in South Korea.” 98Tozer and Leys, “Dust Storms–What Do They Really Cost?” Sources, Costs, and Solutions 17 TABLE 4: MEAN ANNUAL PM2.5, AND TOTAL DEATHS AND LOSSES FROM POLLUTION BY COUNTRY Losses in US Mean Annual Million Dollars % of GDP Country PM2.5 (μg/m3) Total Deaths (PPP-adjusted) Equivalent Algeria 19.26 7,845 9,186 1.84% Bahrain 43.63 188 836 1.47% Egypt 36.41 39,118 33,912 3.85% Iran 31.89 21,680 32,070 2.6% Iraq 32.57 10,372 14,793 2.89% Israel 25.78 2,201 7,639 3.03% Jordan 25.64 1,055 1,083 1.47% Kuwait 49.13 547 3,820 1.44% Lebanon 23.56 1,816 2,808 3.78% Morocco 17.36 7,034 4,158 1.73% Oman 30.35 655 2,725 1.8% Pakistan 46.18 156,191 54,295 6.69% Qatar 38.36 110 1,222 0.44% Saudi Arabia 54.12 6,285 32,038 2.17% Tunisia 16.35 3,792 3,514 3.01% UAE 40.95 900 5,761 1.02% West Bank and Gaza 26.36 1,006 309 1.65% Yemen 36.19 13,442 3,229 3.45% Source: World Bank and IHME, 2016. 18 Sand and Dust Storms in the Middle East and North Africa (MENA) Region PREVENTATIVE ACTIONS, INTERVENTIONS, AND POLICIES AGAINST SAND AND DUST STORMS Investment in prediction technologies, interventions against wind erosion, and regional air pollution policies can significantly reduce the high costs asso- ciated with sand and dust storms. As previously mentioned, about 13 billion USD in GDP are lost every year due to dust storms in the MENA region. Additionally, dust pollution is also linked with many adverse health impacts such as strokes, heart disease, lung cancer, and respiratory diseases like asthma. Preventative and mitigating interventions addressing sand and dust storms are therefore imperative. While dust storms and dust emissions mostly originate from natural sources and are dependent on wind, there are still actions that can be taken to lessen the impacts. Investing in early warning and prediction systems can be extremely beneficial, as it can better prepare economies and significantly lower the dam- ages from sand and dust storms. Additionally, while most dust transport is linked to natural sources, anthropogenic drivers are becoming increasingly threatening too. Government pol- icy addressing these barriers, especially on a transboundary level, should also be prioritized. EARLY WARNING SYSTEMS FOR DUST STORM PREDICTION Early warning systems can reduce many unexpected costs that result from sand and dust storms. Early warning systems at the national and regional levels could prepare people for dust storms and reduce costs. Some major costs include crop losses, adverse health impacts, and infrastructure damage. It gives people time to take cover, seal doors, and vacate streets which prevent car accidents. It would also reduce flight disruption costs, as airlines can activate programs to reschedule or cancel flights before passengers arrive at the airport. Warnings would also give farmers time to bring in livestock and equipment, and allow them to harvest most of a crop if necessary. Sand and dust storms can be predicted by ground-based technologies, space- borne observations, or a combination of both. There are many types of sand and dust storms with respect to spatial and temporal coverage. The physical parameters of a dust storm include its optical depth, concentration, particle size distribution, and land surface Sources, Costs, and Solutions 19 cover below it.99 The main issues in developing a system for dust storm detection and predic- tion include defining data requirements, modeling dust, and designing an effective predic- tion technique. Such systems require data on dust and other environmental changes, which can be obtained in two ways: ground-based observations and spaceborne observations.100 The ground-based observations can be made through lookout towers, video surveillance, and sensory information gathered by radars, lidars, Wireless Sensor Networks (WSNs), etc. The spaceborne observations are typically obtained through satellite imaging, and some- times through unmanned aerial vehicle (UAV), etc.101 However, hybrid approaches make use of both types of observations for better performance. Figure 9 classifies the most commonly used technologies for sand and dust storm detection and prediction. FIGURE 9: TECHNOLOGIES FOR SAND AND DUST STORM MONITORING Sand and dust storm monitoring Ground-based Spaceborne Hybrid observations observations approaches Satellite imaging Lookout towers Satellite imaging + Video surveillance Satellite imaging Unmanned aerial Video surveillance + vehicle Sensory information Satellite imaging Sensory information + Unmanned aerial vehicle Source: Akhlaq et al., 2012. Latest satellite technologies have provided reliable and high-quality data for dust detection. Satellite instruments, particularly the Moderate Resolution Imaging Spect- roradiometer (MODIS), have revolutionized the scientific community’s ability to understand the spatial extent, pathways, and source area of dust storms. Coupling satellite images with Weather Research and Forecasting (WRF) modelling provides a tool for more accurate fore- casting. Other systems such as the Inertial Altitude Reference System (IARS) and the Infra- red Atmospheric Sounding Interferometer (IASI), have the potential to provide good quality dust information as well.102, 103 A reliable Internet connectivity set up to access, download, and transfer data from satellite-borne sensors and terrestrial meteorological stations is key to the success of this time-critical activity. 99El-Askary et al., “Introducing New Approaches for Dust Storms Detection Using Remote Sensing Technology.” 100Ma et al., “New Dust Aerosol Identification Method for Spaceborne Lidar Measurements.” 101Akhlaq, Sheltami, and Mouftah, “A Review of Techniques and Technologies for Sand and Dust Storm Detection.” 102Klüser, Martynenko, and Holzer-Popp, “Thermal Infrared Remote Sensing of Mineral Dust over Land and Ocean.” 103Hilton et al., “Hyperspectral Earth Observation from IASI.” 20 Sand and Dust Storms in the Middle East and North Africa (MENA) Region In addition to weather services and media, warnings can be communicated through text alerts and websites that are quicker and reach larger populations. Early warnings of dust hazards can be communicated through a variety of means, includ- ing media coverage and Short Message Service (SMS) alerts. In South Korea, warnings of yellow dust events transported across the Korean peninsula from China and Mongolia are issued by the Korea Meteorological Administration (KMA) using local media and Short Message Service (SMS) text alerts issued to users who register on their air quality alert web- site. Similarly, the National Weather Service (NWS) in the USA also provides dust storm warnings via SMS.104 The KMA has provided text message alerts since at least 2006, and the USA just began this system in the summer of 2012 as part of a severe weather alert initiative. Although an SMS alert system may not work in developing nations due to poor network connectivity in some areas, it is still a good model for quick and digestible information that can reach a large population. Another example of an early warning platform that updates frequently is a website hosted by the National Centre of Meteorology and Seismology in the United Arab Emirates (UAE), which updates warnings every three hours and provides dust and visibility conditions to the public and local media.105 The World Meteorological Organization (WMO) launched a sand and dust storm warning system that aims to deliver reliable dust storm forecasts through a network of research organizations all over the world. The WMO pro- vides global coordination of monitoring, prediction, and warning systems for sand and dust storms. In 2007, the WMO endorsed the launching of the Sand and Dust Storm Warning Advisory and Assessment System (SDS-WAS). It aims to improve the ability of countries to deliver quick and high-quality sand and dust storm forecasts and knowledge to users through an international partnership of research and operational organizations (Table 5).106 The SDS-WAS works as a network of research, operational centers, and users which is organized through regional nodes.107 Three regional nodes are currently in operation: (i) Northern Africa, Middle East, and Europe, (ii) Asia and Central Pacific, and (iii) Pan-America. The SDS-WAS regional node for Northern Africa, Middle East, and Europe is coordinated by a Regional Centre (RC) set in Barcelona, Spain, and aims to facilitate user access to observa- tional and forecast products and other sources of basic information related to airborne dust. Its web portal (NA-ME-E 2016) provides users with the information needed to monitor dust events and to issue operational predictions and warning advisories related to the dust content in the atmosphere.108 104NOAA, “Mobile Weather Warnings on the Way! National Oceanic and Atmospheric Administration.” 105NCMS, “Warnings. National Centre of Meteorology and Seismology.” 106Terradellas, Nickovic, and Zhang, “Airborne Dust.” 107Nickovic et al., “Sand and Dust Storm Warning Advisory and Assessment System (SDS-WAS) Science and Implementation Plan.” 108Terradellas, Baldasano, and Cuevas Agulló, “Regional Center for Northern Africa, Middle East and Europe of the WMO Sand and Dust Storm Warning Advisory and Assessment Sytem.” Sources, Costs, and Solutions 21 TABLE 5: INSTITUTIONS AND ORGANIZATIONS WITH DUST FORECASTING PROGRAMS Name Location Coverage Japan Meteorological Society Japan Global The Meteorological State Agency of Spain (AEMET) Spain North Africa, Middle and Barcelona Supercomputing Centre East, Europe, Asia Centre for Atmosphere Watch and Services (CAWAS), China East Asia, Central Chinese Meteorological Agency Pacific National Observatory of Athens, University of Greece North Africa, Middle Athens East, Europe, Asia National Research Council Italy North Africa, Middle East, Europe, Asia South east European Climate Change Center Serbia North Africa, Middle East, Europe, Asia Finnish Meteorological Institute Finland North Africa, Middle East, Europe, Asia Korean Meteorological Administration Korea East Asia University of Tel Aviv Israel North Africa, Middle East, Europe Egyptian Meteorological Authority Egypt North Africa, Middle East, Europe, Asia Naval Research Laboratory, National Aeronautics USA Global and space Administration, National Centers for Environmental Prediction Research Institute for Applied Mechanics, Kyushu Japan East Asia, Central University in cooperation with the National Institute Pacific for Environmental Studies (NIES) Laboratoire de Meteorologie Dynamique France Africa, Europe, Atlantic, Central Asia European Center for Medium range Weather UK Global Forecasting (ECMWF), Met Office World Meteorological Organization (WMO) Switzerland Global network of regional models Source: WMO, 2011109, UNEP, WMO, UNCCD, 2016110 TECHNICAL INTERVENTIONS Agricultural practices, such as residue management, tillage, shelterbelts, and agroforestry, are effective against wind erosion and decreasing dust transport. To prevent dust transport and dust storms, soil exposure to wind must be managed. This can be done by protecting the soil with live or dead vegetation or minimizing the time and area of the soil that has little cover. Additionally, adding windbreaks and adopting agroforestry in agricultural land can also reduce wind erosion Cropping, residue management, shelterbelts, and reduced tillage practices are especially helpful for preventing wind erosion. Reduced or no tillage, and other practices that minimize soil disturbance and maximize residue on the soil are included under ‘conservation agriculture’. Such practices also reduce the time that 109WMO, “Organizations Delivering SDS Forecasts. World Meteorology Organization.” 110 UNEP, WMO, UNCCD (2016). Global Assessment of Sand and Dust Storms. United Nations Environment Programme, Nairobi 22 Sand and Dust Storms in the Middle East and North Africa (MENA) Region the soil is unprotected in dry seasons. Other good management practices include use of qual- ity planting material, optimal plant density, appropriate soil and crop nutrient management, and adequate pest and disease control. Maintaining crop residue successfully protects the soil from erosion. Maintain- ing enough vegetative cover is often referred to as the “cardinal rule” for controlling wind erosion.111 Crop residues as a cover help stabilize the soil by reducing soil water loss and erosivity. For example, Michels et al. (1995) showed that covering the soil with 2,000 kg per ha of millet residues provides enough protection from sand storms.112 Mulching—the practice of leaving some residual crop material such as leaves, stalks, and roots, on or near the surface—is a commonly used agronomic technique. It is successful in reducing erosion and in reducing the loss of water from fields by decreasing evaporation. Research shows that vertical residues are also very effective for controlling soil loss during wind, as they can trap more snow than horizontal stems for instance.113, 114 Windbreaks such as shelterbelts reduce wind speed and provide other agricul- tural benefits as well. Windbreaks are structures that reduce wind speed and are com- monly associated with a natural vegetative barrier against wind.115 These types of barriers can inhibit wind erosion by reducing the travel distance of wind across a field. Fences or walls placed at right angles to erosive winds can reduce wind erosion, or windbreaks may be cre- ated from living plants such as trees or bushes, in which case they are known as shelterbelts. Reductions in wind velocity are achieved both upwind, for 2–5 times the height of the wind- break, and downwind, extending 10–30 times of the windbreak height.116 Shelterbelts pro- duce many benefits for farmers, such as decreased soil erosion, increased crop yields, reduced livestock stress, control of drifting snow, building maintenance, and energy savings.117 Other benefits include increased soil and air temperatures, reduced pest and disease problems, and an extended growing season in sheltered areas. Agroforestry, along with providing other ecosystem services, offers wind- breaks in agriculture land.118 Agroforestry includes linear arrangements of trees and shrubs around fields and homesteads, along roadsides, on soil conservation contours within fields, and in riparian areas. Scattered trees are especially important for protecting croplands in dryland areas. Poor farmers who are mostly concerned with crop yields will not prioritize wind erosion as much.119 However, introducing agroforestry will provide economic incen- tives as well as protection against wind erosion. NATIONAL AND REGIONAL GOVERNMENT POLICIES Besides investing in early warning systems, governments all over the world are designing policies to mitigate the impact of sand and dust storms, both at national and regional levels. Devastating impacts from sand and dust storms in the 111Skidmore, “Wind Erosion Climatic Erosivity.” 112Michels et al., “Wind and Windblown Sand Damage to Pearl Millet.” 113Bilbro and Fryrear, “Wind Erosion Losses as Related to Plant Silhouette and Soil Cover.” 114Nielsen, Hinkle, and Lyon, “Wind Velocity, Snow and Soil Water Measurements in Sunflower Residues.” 115Rosenberg, Blad, and Verma, Microclimate. 116Cornelis and Gabriels, “Optimal Windbreak Design for Wind-Erosion Control.” 117Forman and Baudry, “Hedgerows and Hedgerow Networks in Landscape Ecology.” 118Young, “Agroforestry for Soil Conservation.” 119Sterk, “Causes, Consequences and Control of Wind Erosion in Sahelian Africa.” Sources, Costs, and Solutions 23 BOX 3: UN-INTERAGENCY RESPONSE TO SDS The United Nations Coalition on Combatting SDS was launched at COP 14. The UN Coa- lition was established in response to the United Nations General Assembly resolution 72/225 in 2017 through the efforts made by UNEP. Currently 15 members of the coalition include: UNEP, WMO, UNCCD, UNITAR, ICAO, UNDP, UN-Habitat, WHO, ESCAP, ESCWA, IUCN, FAO, World Bank, ITU, and UNECE. The key objectives of the coalition include: » Prepare a global response to SDS, including a strategy and an action plan, which could result in the development of a United Nations system-wide approach to addressing SDS. Identifying entry points to support SDS-affected countries and regions in the implementa- tion of cross-sectoral, and transboundary risk reduction and response measures for SDS. » Provide a forum for engaging with partners and enhancing dialogue and collaboration among affected countries and the UN system agencies at global, regional, and subre- gional levels. » Provide a common platform for exchange of knowledge, data, information, and techni- cal expertise and resources for strengthening preparedness measures and strategies for risk reduction, consolidated policy, innovative solutions, advocacy and capacity building efforts, and fund-raising initiatives. » Identify, mobilize, and facilitate access to financial resources for joint responses to sand and dust storms, including through new and innovative resources and mechanisms. Four main cross-cutting work areas will be addressed by the coalition: » Facilitation of information exchange among stakeholders (e.g., data collection, knowl- edge sharing, and innovative solutions) » Capacity building and training » Mobilizing resources and fund-raising initiatives » Advocacy and awareness raising For more information see: https://unemg.org/our-work/emerging-issues/sand-and-dust- storms/ Source: UN Environment Management Group, 2019. Americas, MENA region, and East Asia have encouraged governments to enforce many large-scale initiatives and plans. In many instances, these initiatives also tackle land degra- dation, terrestrial biodiversity, and climate change mitigation. However, policies designed to mitigate the wider impacts of sand and dust storms, including many that are transboundary, are geographically patchy and have a much shorter history. Regional and international coop- eration among countries will lead to a greater understanding of the transportation paths of dust storms, particle content, and their impacts. A new United Nations Coalition on Com- batting Sand and Dust Storms has recently been launched with the goal of raising aware- ness, capacity building, and mobilizing resources to respond to SDS (Box 3). Eventually, regional action will also lead to reduced occurrence of dust storms. Recent years have seen some regional air pollution policies emerge, but more collaboration is needed and should be sustained. Large-scale disasters like the dust bowl in the USA in the 1930s motivated sev- eral government initiatives and policies against sand and dust storms. Even 24 Sand and Dust Storms in the Middle East and North Africa (MENA) Region though the Dust Bowl in the 1930s in the USA had devasting impacts all around, one silver lining was the government’s increased participation in soil conservations and land manage- ment issues. For instance, the Soil Conservation Service (SCS), created in 1935, identified areas in need of remediation using aerial photography surveys and detailed soil maps.120 They acquired abandoned lands, which are dust storm sources, and used them for demon- stration projects on terracing and contour plowing. Other government agencies provided subsidies to encourage improved plowing methods and funded planting of shelterbelts on many private lands. Similar interventions by the government also occurred in Canada. Some farmland was converted to rangelands, subsidies were offered to families willing to aban- don farms in dry areas, and farmers were encouraged to build shelterbelts and adopt soil conservation.121 In addition to government initiatives after a dust storm disaster, several coun- tries have developed government policies to mitigate the impacts of sand and dust storms. In China, one of the most ambitious projects to combat desertification and control dust storms is an afforestation project called the Three Norths Forest Shelterbelt pro- gram or the Great Green Wall, which is to be completed by 2050.122 Other projects in China include the Grain-for-Green program, which is designed to convert cropland to forest and grassland and the Beijing-Tianjin Sand Source Control program, which includes various interventions that conserve cropland and grazing lands and prevent erosion.123, 124 China also has a National Action Plan to implement the United Nations Convention to Combat Desertification (UNCCD), drawn up in 1996 and revised in 2003, and is the first country to establish a national desertification monitoring initiative as a follow-up action. In a similar initiative to the Great Green Wall of China, the African Union, with the support of the World Bank, is establishing a Great Green Wall of trees and shrubs along the southern edge of the Sahara Desert. The project aims to reforest 15 million hectares along a 15 km-wide, 7,775 km-long belt, from Dakar to Djibouti. Northeast Asia and West Asia have developed Regional Action Plans on sand and dust storms, which aim to monitor dust storms and invest in mitigation strategies. The Regional Master Plan for the Prevention and Control of Dust and Sand- storms in Northeast Asia is a project involving the governments of China, Japan, Mongo- lia, and South Korea.125 It was jointly initiated and conducted by the Asian Development Bank (ADB), the UNCCD, the United Nations Economic and Social Commission for Asia and Pacific (UNESCAP), and United Nations Environment Programme (UNEP). The plan’s aims are two-fold—to establish a regional monitoring, forecasting, and early warning net- work for dust storms in Northeast Asia and to invest in strengthening mitigation measures against root causes of dust storms in the regional source areas. Another regional master plan is the West Asia Regional Master Plan to Combat Sand and Dust Storms, which is coordi- nated by UNEP and WMO Regional Offices for West Asia Iran.126, 127 The plan includes Bahrain, Iran, Iraq, Jordan, Kuwait, Oman, Qatar, Saudi Arabia, Syria, Turkey, and the UAE. There are also several other global and regional air pollution agreements and policies that are addressing dust pollution and dust storm monitoring (Box 4). 120McLeman et al., “What We Learned from the Dust Bowl: Lessons in Science, Policy, and Adaptation.” 121Marchildon et al., “Drought and Institutional Adaptation in the Great Plains of Alberta and Saskatchewan, 1914–1939.” 122Wang et al., “Has the Three Norths Forest Shelterbelt Program Solved the Desertification and Dust Storm Problems in Arid and Semiarid China?” 123Lei, Shangguan, and Rui, “Effects of the Grain-for-Green Program on Soil Erosion in China.” 124Dalintai, Yanbo, and Jianjun, “The Eurasian Steppe.” 125Diallo, “United Nations Convention to Combat Desertification (UNCCD).” 126Cuevas, Establishing a WMO Sand and Dust Storm Warning Advisory and Assessment System Regional Node for West Asia. 127UNEP, “West Asia Regional Master Plan to Combat Sand and Dust Storms. United Nations Environment Programme.” Sources, Costs, and Solutions 25 BOX 4: REGIONAL AIR POLLUTION POLICIES The Malé Declaration on Control and Prevention of Air Pollution and Its Likely Transbound- ary Effects for South Asia is an intergovernmental agreement to tackle regional air pollution problems, established in 1998 by the South Asian countries at a meeting of the South Asia Cooperative Environment Programme (SACEP) Governing Council. Regional Air Pollution in Developing Countries (RAPIDC) aims to facilitate the development of agreements/protocols to implement measures which prevent and control air pollution. The Air Pollution Information Network for Africa (APINA) was formed in 1997 and acts as a link between different networks and programs on air pollution in Africa. The governments of the ten ASEAN (Association of Southeast Asian Nations) member coun- tries signed the ASEAN Agreement on Transboundary Haze Pollution in 2002. The agreement is the first regional agreement that binds a group of countries to tackle transboundary haze pollution resulting from land and forest fires. China—South Korea—Japan joint sand and dust storm cooperation within the Asia Node of SDSWAS. China—South Korea—Mongolia joint sand and dust storm monitoring. Mongolia and Kazakhstan are associated partners of the Asia Node of SDS-WAS. The Convention on Long-Range Transboundary Air Pollution (CLRTAP) is intended to pro- tect humans and the environment against air pollution and to gradually reduce and prevent air pollution, including long-range transboundary air pollution exchange. Source: Shepherd et al., 2016. 26 Sand and Dust Storms in the Middle East and North Africa (MENA) Region REFERENCES Akhlaq, Muhammad, Tarek R. Sheltami, and Hussein T. Mouftah. “A Review of Tech- niques and Technologies for Sand and Dust Storm Detection.” Reviews in Environmental Science and Bio/Technology 11, no. 3 (2012): 305–322. Al Frayh, A. R., Z. Shakoor, M. O. Gad ElRab, and S. M. Hasnain. “Increased Prevalence of Asthma in Saudi Arabia.” Annals of Allergy, Asthma & Immunology 86, no. 3 (2001): 292–296. Anuforom, A. C., L. E. Akeh, P. N. Okeke, and F. E. Opara. “Inter-Annual Variability and Long-Term Trend of UV-Absorbing Aerosols during Harmattan Season in Sub-­ Saharan West Africa.” Atmospheric Environment 41, no. 7 (2007): 1550–1559. Assessment, Millennium Ecosystem. Ecosystems and Human Well-Being. Vol. 5. Island Press, Washington, DC (2005). Aumont, Olivier, Laurent Bopp, and M. Schulz. “What Does Temporal Variability in Aeo- lian Dust Deposition Contribute to Sea-Surface Iron and Chlorophyll Distributions?” Geophysical Research Letters 35, no. 7 (2008). Behzad, Hayedeh, Katsuhiko Mineta, and Takashi Gojobori. “Global Ramifications of Dust and Sandstorm Microbiota.” Genome Biology and Evolution 10, no. 8 (2018): 1970–1987. Bener, A., Y. M. Abdulrazzaq, J. Al-Mutawwa, and P. Debuse. “Genetic and Environmental Factors Associated with Asthma.” Human Biology, 1996, 405–414. Bilbro, J. D., and D. W. Fryrear. “Wind Erosion Losses as Related to Plant Silhouette and Soil Cover.” Agronomy Journal 86, no. 3 (1994): 550–553. Bullard, Joanna E., Sandy P. Harrison, Matthew C. Baddock, Nick Drake, Thomas E. Gill, Grant McTainsh, and Youbin Sun. “Preferential Dust Sources: A Geomorphological Classification Designed for Use in Global Dust-Cycle Models.” Journal of Geophysical Research: Earth Surface 116, no. F4 (2011). Cao, Hui, Farshad Amiraslani, Jian Liu, and Na Zhou. “Identification of Dust Storm Source Areas in West Asia Using Multiple Environmental Datasets.” Science of the Total Environ- ment 502 (2015): 224–235. Chadwick, Oliver A., Louis A. Derry, Peter M. Vitousek, Barry J. Huebert, and Lars O. Hedin. “Changing Sources of Nutrients during Four Million Years of Ecosystem Devel- opment.” Nature 397, no. 6719 (1999): 491. Chan, Chang-Chuan, Kai-Jen Chuang, Wen-Jone Chen, Wei-Tien Chang, Chung-Te Lee, and Chi-Ming Peng. “Increasing Cardiopulmonary Emergency Visits by Long-Range Sources, Costs, and Solutions 27 Transported Asian Dust Storms in Taiwan.” Environmental Research 106, no. 3 (2008): 393–400. Chang, Chih-Ching, I.-Ming Lee, Shang-Shyue Tsai, and Chun-Yuh Yang. “Correlation of Asian Dust Storm Events with Daily Clinic Visits for Allergic Rhinitis in Taipei, Taiwan.” Journal of Toxicology and Environmental Health, Part A 69, no. 3 (2006): 229–235. Cheng, Ming-Fen, Shu-Chen Ho, Hui-Fen Chiu, Trong-Neng Wu, Pei-Shih Chen, and Chun- Yuh Yang. “Consequences of Exposure to Asian Dust Storm Events on Daily Pneumonia Hospital Admissions in Taipei, Taiwan.” Journal of Toxicology and Environmental Health, Part A 71, no. 19 (2008): 1295–1299. Choobari, O. Alizadeh, P. Zawar-Reza, and A. Sturman. “The Global Distribution of Min- eral Dust and Its Impacts on the Climate System: A Review.” Atmospheric Research 138 (2014): 152–165. Cornelis, W. M., and Donald Gabriels. “Optimal Windbreak Design for Wind-Erosion Con- trol.” Journal of Arid Environments 61, no. 2 (2005): 315–332. Cowie, Sophie M., Peter Knippertz, and John H. Marsham. “Are Vegetation-Related Rough- ness Changes the Cause of the Recent Decrease in Dust Emission from the Sahel?” Geo- physical Research Letters 40, no. 9 (2013): 1868–1872. Cuevas, E. Establishing a WMO Sand and Dust Storm Warning Advisory and Assessment System Regional Node for West Asia: Current Capabilities and Needs. WMO (2013). Dalintai, B., Li Yanbo, and Cao Jianjun. “The Eurasian Steppe: History of Utilization and Policies on the Rangeland.” Restoring Community Connections to the Land (2011): 51–68. De Longueville, Florence, Yvon-Carmen Hountondji, Sabine Henry, and Pierre Ozer. “What Do We Know about Effects of Desert Dust on Air Quality and Human Health in West Africa Compared to Other Regions?” Science of the Total Environment 409, no. 1 (2010): 1–8. Delangizan, Sohrab, and Zainab Jafari Motlagh. “Dust Phenomenon Affects on Cardiovas- cular and Respiratory Hospitalizations and Mortality: A Case Study in Kermanshah, during March–September 2010–2011.” Iranian Journal of Health and Environment 6, no. 1 (2013): 65–76. Derbyshire, Edward. “Natural Minerogenic Dust and Human Health.” AMBIO: A Journal of the Human Environment 36, no. 1 (2007): 73–78. Diallo, Hama Arba. “United Nations Convention to Combat Desertification (UNCCD).” In The Future of Drylands, 13–16. Springer (2008). Draxler, Roland R., Dale A. Gillette, Jeffrey S. Kirkpatrick, and Jack Heller. “Estimating PM10 Air Concentrations from Dust Storms in Iraq, Kuwait and Saudi Arabia.” Atmo- spheric Environment 35, no. 25 (2001): 4315–4330. El-Askary, Hesham, Menas Kafatos, Xue Liu, and Tarek El-Ghazawi. “Introducing New Approaches for Dust Storms Detection Using Remote Sensing Technology.” In IGARSS 2003. 2003 IEEE International Geoscience and Remote Sensing Symposium. Proceed- ings (IEEE Cat. No. 03CH37477), 4:2439–2441. IEEE (2003). Eswaran, H., R. Lal, and P. F. Reich. “Land Degradation: An Overview. Responses to Land Degradation.” In Proc. 2nd International Conference on Land Degradation and Desert- ification. Khon Kaen, Thailand, edited by E. Bridges, I. Hannam, L. Oldeman, F. Pen- ning de Vries, S. Scherr, and S. Sompatpanit. New Delhi: Oxford Press (2001). Evan, Amato T., Jason Dunion, Jonathan A. Foley, Andrew K. Heidinger, and Christo- pher S. Velden. New Evidence for a Relationship between Atlantic Tropical Cyclone Activity and African Dust Outbreaks (2006). http://citeseerx.ist.psu.edu/viewdoc/ summary?doi=10.1.1.526.7037 28 Sand and Dust Storms in the Middle East and North Africa (MENA) Region Fairlie, T. Duncan, Daniel J. Jacob, and Rokjin J. Park. “The Impact of Transpacific Trans- port of Mineral Dust in the United States.” Atmospheric Environment 41, no. 6 (2007): 1251–1266. Forman, Richard T. T., and Jacques Baudry. “Hedgerows and Hedgerow Networks in Land- scape Ecology.” Environmental Management 8, no. 6 (1984): 495–510. Furman, Haim Kutiel Hadar. “Dust Storms in the Middle East: Sources of Origin and Their Temporal Characteristics.” Indoor and Built Environment 12, no. 6 (2003): 419–426. Gakidou, Emmanuela, Ashkan Afshin, Amanuel Alemu Abajobir, Kalkidan Hassen Abate, Cristiana Abbafati, Kaja M. Abbas, Foad Abd-Allah, Abdishakur M. Abdulle, Semaw Ferede Abera, and Victor Aboyans. “Global, Regional, and National Comparative Risk Assessment of 84 Behavioural, Environmental and Occupational, and Metabolic Risks or Clusters of Risks, 1990–2016: A Systematic Analysis for the Global Burden of Dis- ease Study 2016.” The Lancet 390, no. 10100 (2017): 1345–1422. Garrison, Virginia H., Eugene A. Shinn, William T. Foreman, Dale W. Griffin, Charles W. Holmes, Christina A. Kellogg, Michael S. Majewski, Laurie L. Richardson, Kim B. Ritchie, and Garriet W. Smith. “African and Asian Dust: From Desert Soils to Coral Reefs.” BioScience 53, no. 5 (2003): 469–480. Gill, Thomas E. “Eolian Sediments Generated by Anthropogenic Disturbance of Playas: Human Impacts on the Geomorphic System and Geomorphic Impacts on the Human System.” Geomorphology 17, no. 1–3 (1996): 207–228. Ginoux, Paul, Joseph M. Prospero, Thomas E. Gill, N. Christina Hsu, and Ming Zhao. “Global-Scale Attribution of Anthropogenic and Natural Dust Sources and Their Emis- sion Rates Based on MODIS Deep Blue Aerosol Products.” Reviews of Geophysics 50, no. 3 (2012). Global Asthma Network. “Global Asthma Report” (2014). https://www.who.int/ quantifying_ehimpacts/national/countryprofile/intro/en/ Goudie, Andrew S. “Dust Storms: Recent Developments.” Journal of Environmental Manage- ment 90, no. 1 (2009): 89–94. Grineski, Sara E., Joan G. Staniswalis, Priyangi Bulathsinhala, Yanlei Peng, and Thomas E. Gill. “Hospital Admissions for Asthma and Acute Bronchitis in El Paso, Texas: Do Age, Sex, and Insurance Status Modify the Effects of Dust and Low Wind Events?” Environ- mental Research 111, no. 8 (2011): 1148–1155. Grousset, Francis E., Paul Ginoux, Aloys Bory, and Pierre E. Biscaye. “Case Study of a Chi- nese Dust Plume Reaching the French Alps.” Geophysical Research Letters 30, no. 6 (2003). Gwack, J., S. S. Hwang, H. J. Kwon, Y. C. Hong, S. H. Cho, Y. Yoo, J. H. Yoo, and Y. R. Ko. “Effects of Asian Dust Events on Diurnal Variation of Peak Expiratory Flow Rate in Children with Bronchial Asthma and Healthy Children.” Epidemiology 16, no. 5 (2005): S109–S110. Gyan, K., W. Henry, S. Lacaille, A. Laloo, C. Lamsee-Ebanks, S. McKay, R. M. Antoine, and Michele Anne Monteil. “African Dust Clouds Are Associated with Increased Pae- diatric Asthma Accident and Emergency Admissions on the Caribbean Island of Trini- dad.” International Journal of Biometeorology 49, no. 6 (2005): 371–376. Hamidi, Mehdi, Mohammad Reza Kavianpour, and Yaping Shao. “Synoptic Analysis of Dust Storms in the Middle East.” Asia-Pacific Journal of Atmospheric Sciences 49, no.  3 (2013): 279–286. Highwood, Ellie, and Claire Ryder. “Radiative Effects of Dust.” Springer (2014). http:// centaur.reading.ac.uk/70210/ Sources, Costs, and Solutions 29 Hilton, Fiona, Raymond Armante, Thomas August, Chris Barnet, Aurelie Bouchard, Claude Camy-Peyret, Virginie Capelle, Lieven Clarisse, Cathy Clerbaux, and Pierre-Francois Coheur. “Hyperspectral Earth Observation from IASI: Five Years of Accomplishments.” Bulletin of the American Meteorological Society 93, no. 3 (2012): 347–370. Hong, Yun-Chul, Xiao-Chuan Pan, Su-Young Kim, Kwangsik Park, Eun-Jung Park, Xiaobin Jin, Seung-Muk Yi, Yoon-Hee Kim, Choong-Hee Park, and Sanghwan Song. “Asian Dust Storm and Pulmonary Function of School Children in Seoul.” Science of the Total Environment 408, no. 4 (2010): 754–759. Jeong, Dai-Yeun. “Socio-Economic Costs from Yellow Dust Damages in South Korea.” Korean Social Science Journal 35 (2008): 1–29. Jickells, T. D., S. Dorling, W. G. Deuser, T. M. Church, R. Arimoto, and J. M. Prospero. “Air-Borne Dust Fluxes to a Deep Water Sediment Trap in the Sargasso Sea.” Global Biogeochemical Cycles 12, no. 2 (1998): 311–320. Kanatani, Kumiko T., Isao Ito, Wael K. Al-Delaimy, Yuichi Adachi, William C. Mathews, and Joe W. Ramsdell. “Desert Dust Exposure Is Associated with Increased Risk of Asthma Hospitalization in Children.” American Journal of Respiratory and Critical Care Med- icine 182, no. 12 (2010): 1475–1481. Klüser, Lars, D. Martynenko, and T. Holzer-Popp. “Thermal Infrared Remote Sensing of Mineral Dust over Land and Ocean: A Spectral SVD Based Retrieval Approach for IASI.” Atmospheric Measurement Techniques 4, no. 5 (2011): 757–773. Knippertz, Peter, and Jan-Berend W. Stuut. “Mineral Dust.” Springer Dordrecht Heidel- berg, New York, London, Doi 10 (2014): 978–994. Lee, J. A., T. E. Gill, and K. R. Mulligan. “The 1930s Dust Bowl: The Relative Roles of Peo- ple and the Physical Environment.” In Geological Society of America Abstracts and Programmes, 31:A11 (1999). Lee, Joong-Woo, and Ki-Kwang Lee. “Effects of Asian Dust Events on Daily Asthma Patients in Seoul, Korea.” Meteorological Applications 21, no. 2 (2014): 202–209. Lei, Deng, Zhou-Ping Shangguan, and L. I. Rui. “Effects of the Grain-for-Green Program on Soil Erosion in China.” International Journal of Sediment Research 27, no. 1 (2012): 120–127. Liu, Zhaoyan, Ali Omar, Mark Vaughan, Johnathan Hair, Chieko Kittaka, Yongxiang Hu, Kathleen Powell, Charles Trepte, David Winker, and Chris Hostetler. “CALIPSO Lidar Observations of the Optical Properties of Saharan Dust: A Case Study of Long-Range Transport.” Journal of Geophysical Research: Atmospheres 113, no. D7 (2008). Ma, Yingying, Wei Gong, Pucai Wang, and Xiuqing Hu. “New Dust Aerosol Identifica- tion Method for Spaceborne Lidar Measurements.” Journal of Quantitative Spectroscopy and Radiative Transfer 112, no. 2 (2011): 338–345. Mahowald, Natalie M., Silvia Kloster, S. Engelstaedter, J. Keith Moore, S. Mukhopadhyay, Joseph R. McConnell, S. Albani, Scott C. Doney, A. Bhattacharya, and M. A. J. Curran. “Observed 20th Century Desert Dust Variability: Impact on Climate and Biogeochem- istry.” Atmospheric Chemistry and Physics 10, no. 22 (2010): 10875–10893. Marchildon, Gregory P., Suren Kulshreshtha, Elaine Wheaton, and Dave Sauchyn. “Drought and Institutional Adaptation in the Great Plains of Alberta and Saskatchewan, 1914–1939.” Natural Hazards 45, no. 3 (2008): 391–411. McLeman, Robert A., Juliette Dupre, Lea Berrang Ford, James Ford, Konrad Gajewski, and Gregory Marchildon. “What We Learned from the Dust Bowl: Lessons in Science, Pol- icy, and Adaptation.” Population and Environment 35, no. 4 (2014): 417–440. 30 Sand and Dust Storms in the Middle East and North Africa (MENA) Region Meng, Ziqiang, and Bin Lu. “Dust Events as a Risk Factor for Daily Hospitalization for Respiratory and Cardiovascular Diseases in Minqin, China.” Atmospheric Environment 41, no. 33 (2007): 7048–7058. Michels, Karlheinz, Dean V. Armbrust, Bruce E. Allison, and Mannava VK Sivakumar. “Wind and Windblown Sand Damage to Pearl Millet.” Agronomy Journal 87, no. 4 (1995): 620–626. Middleton, N., and A. S. Goudie. Desert Dust in the Global System. Springer (2006). Middleton, N. J. “Dust Storms in the Middle East.” Journal of Arid Environments 10, no. 2 (1986): 83–96. Moridnejad, Ali, Neamat Karimi, and Parisa A. Ariya. “A New Inventory for Middle East Dust Source Points.” Environmental Monitoring and Assessment 187, no. 9 (2015): 582. Muhs, Daniel R., Joseph M. Prospero, Matthew Baddock, and Thomas E. Gill. “Identifying Sources of Aeolian Mineral Dust: Present and Past.” Springer (2014). http://hdl.handle .net/10072/140539 Mulitza, Stefan, David Heslop, Daniela Pittauerova, Helmut W. Fischer, Inka Meyer, Jan-Berend Stuut, Matthias Zabel, Gesine Mollenhauer, James A. Collins, and Henning Kuhnert. “Increase in African Dust Flux at the Onset of Commercial Agriculture in the Sahel Region.” Nature 466, no. 7303 (2010): 226. NCMS. “Warnings. National Centre of Meteorology and Seismology” (2013). http://www .ncms.ae/warning/warning3.htm Nickovic, Slodoban, Emilio Cuevas Agulló, José María Baldasano, Enric Terradellas, Tetsuo Nakazawa, and Alexander Baklanov. “Sand and Dust Storm Warning Advisory and Assessment System (SDS-WAS) Science and Implementation Plan: 2015–2020” (2015). Nielsen, D. C., S. E. Hinkle, and D. J. Lyon. “Wind Velocity, Snow and Soil Water Measure- ments in Sunflower Residues.” Great Plains Agric. Council Bull 150 (1994): 93–100. NOAA. “Mobile Weather Warnings on the Way! National Oceanic and Atmospheric Administration” (2012). http://www.noaa.gov/features/03_protecting/wireless_ emergency_alerts.html Noinaj, Nicholas, Susan K. Buchanan, and Cynthia Nau Cornelissen. “The Transferrin– Iron Import System from Pathogenic N Eisseria Species.” Molecular Microbiology 86, no. 2 (2012): 246–257. Oerlemans, J., R.H. Giesen, and M.R. van den Broeke. “Retreating alpine glaciers: increased melt rates due to accumulation of dust (Vadret da Morteratsch, Switzerland)” (2009). http://dspace.library.uu.nl:8080/handle/1874/43789 Parolari, Anthony J., Dan Li, Elie Bou-Zeid, Gabriel G. Katul, and Shmuel Assouline. “Cli- mate, Not Conflict, Explains Extreme Middle East Dust Storm.” Environmental Research Letters 11, no. 11 (2016): 114013. Perdue, Philip, Kirk K. Kazarian, Charles Odeyale, Jerry Quance, Isaac Hayward, and Taffy Williams. “The Surgical Significance of Persian Gulf Sand.” Military Medicine 157, no. 7 (1992): 375–377. Perez, Laura, Aurelio Tobias, Xavier Querol, Nino Künzli, Jorge Pey, Andrés Alastuey, Mar Viana, Natalia Valero, Manuel González-Cabré, and Jordi Sunyer. “Coarse Particles from Saharan Dust and Daily Mortality.” Epidemiology 19, no. 6 (2008): 800–807. Prospero, Joseph M. “Assessing the Impact of Advected African Dust on Air Quality and Health in the Eastern United States.” Human and Ecological Risk Assessment: An International Journal 5, no. 3 (1999): 471–479. Sources, Costs, and Solutions 31 ———. “Long-Range Transport of Mineral Dust in the Global Atmosphere: Impact of African Dust on the Environment of the Southeastern United States.” Proceedings of the National Academy of Sciences 96, no. 7 (1999): 3396–3403. Prospero, Joseph M., Paul Ginoux, Omar Torres, Sharon E. Nicholson, and Thomas E. Gill. “Environmental Characterization of Global Sources of Atmospheric Soil Dust Identi- fied with the Nimbus 7 Total Ozone Mapping Spectrometer (TOMS) Absorbing Aerosol Product.” Reviews of Geophysics 40, no. 1 (2002): 2–31. Prospero, Joseph M., and Olga L. Mayol-Bracero. “Understanding the Transport and Impact of African Dust on the Caribbean Basin.” Bulletin of the American Meteorological Society 94, no. 9 (2013): 1329–1337. Ravi, Sujith, Paolo D’Odorico, David D. Breshears, Jason P. Field, Andrew S. Goudie, Travis E. Huxman, Junran Li, et al. “Aeolian Processes and the Biosphere” (2011). http://hdl .handle.net/2160/11813 Ridley, D. A., C. L. Heald, and J. M. Prospero. “What Controls the Recent Changes in African Mineral Dust Aerosol across the Atlantic?” Atmospheric Chemistry and Physics 14, no. 11 (2014): 5735–5747. Rosenberg, Norman J., Blaine L. Blad, and Shashi B. Verma. Microclimate: The Biological Environment. John Wiley & Sons (1983). Samoli, E., P. T. Nastos, A. G. Paliatsos, K. Katsouyanni, and K. N. Priftis. “Acute Effects of Air Pollution on Pediatric Asthma Exacerbation:Evidence of Association and Effect Modification” (2011). https://kclpure.kcl.ac.uk/portal/en/publications/acute-­ effects-of-air-pollution-on-pediatric-asthma-exacerbation(713c2dec-d523-41db-8550- 5231325c5ec1).html Samoli, Evangelia, Evgenia Kougea, Pavlos Kassomenos, Antonis Analitis, and Klea Kat- souyanni. “Does the Presence of Desert Dust Modify the Effect of PM10 on Mortality in Athens, Greece?” Science of the Total Environment 409, no. 11 (2011): 2049–2054. Sassen, Kenneth, Paul J. DeMott, Joseph M. Prospero, and Michael R. Poellot. “Saharan Dust Storms and Indirect Aerosol Effects on Clouds: CRYSTAL-FACE Results.” Geo- physical Research Letters 30, no. 12 (2003). Shao, Yaping, Martina Klose, and Karl-Heinz Wyrwoll. “Recent Global Dust Trend and Connections to Climate Forcing.” Journal of Geophysical Research: Atmospheres 118, no. 19 (2013): 11–107. Shi, P.J., P. Yan, Y. Yuan, and M. A. Nearing. “Wind Erosion Research in China: Past, Pres- ent and Future.” Progress in Physical Geography 23, 3 (2004): 366–386. Singh, Ramesh P., Anup K. Prasad, Vinay K. Kayetha, and Menas Kafatos. “Enhancement of Oceanic Parameters Associated with Dust Storms Using Satellite Data.” Journal of Geophysical Research: Oceans 113, no. C11 (2008). Skidmore, E. L. “Wind Erosion Climatic Erosivity.” Climatic Change 9, no. 1–2 (1986): 195–208. Sterk, G. “Causes, Consequences and Control of Wind Erosion in Sahelian Africa: A Review.” Land Degradation & Development 14, no. 1 (2003): 95–108. Tam, Wilson W. S., Tze Wai Wong, Andromeda H. S. Wong, and David S. C. Hui. “Effect of Dust Storm Events on Daily Emergency Admissions for Respiratory Diseases.” Respi- rology 17, no. 1 (2012): 143–148. Tanaka, Taichu Y., and Masaru Chiba. “A Numerical Study of the Contributions of Dust Source Regions to the Global Dust Budget.” Global and Planetary Change 52, no. 1–4 (2006): 88–104. 32 Sand and Dust Storms in the Middle East and North Africa (MENA) Region Teather, Kevin, Natacha Hogan, Kim Critchley, Mark Gibson, Susanne Craig, and Janet Hill. Examining the Links between Air Quality, Climate Change and Respiratory Health in Qatar. Bloomsbury Qatar Foundation Journals (2013). http://www.qscience .com/doi/10.5339/avi.2013.9 Terradellas, Enric, José María Baldasano, and Emilio Cuevas Agulló. “Regional Center for Northern Africa, Middle East and Europe of the WMO Sand and Dust Storm Warning Advisory and Assessment Sytem: Activity Report 2010–2012” (2014). Terradellas, Enric, Slobodan Nickovic, and Xiao Ye Zhang. “Airborne Dust: A Hazard to Human Health, Environment and Society.” Boletín-Organización Meteorológica Mundial 64, no. 2 (2015): 44–48. Thalib, Lukman, and Abdullah Al-Taiar. “Dust Storms and the Risk of Asthma Admissions to Hospitals in Kuwait.” Science of the Total Environment 433 (2012): 347–351. Thomas, David S. G. Arid Zone Geomorphology: Process, Form and Change in Drylands. John Wiley & Sons (2011). Tozer, P., and J. Leys. “Dust Storms—What Do They Really Cost?” The Rangeland Journal 35, no. 2 (2013): 131–142. Trigo, Isabel F., Trevor D. Davies, and Grant R. Bigg. “Objective Climatology of Cyclones in the Mediterranean Region.” Journal of Climate 12, no. 6 (1999): 1685–1696. UN Environment Management Group. “United Nations Coalition on Combatting Sand and Dust Storms (SDS)—Terms of Reference.” May 2019. UNEP. “Global Environment Outlook GEO–4: Environment for Development.” United Nations Environment Programme, Nairobi, 2007. UNEP. “West Asia Regional Master Plan to Combat Sand and Dust Storms. United Nations Environment Programme” (2014). UNEP. “Sand and Dust Storms,” United Nations Environment Assembly, UNEA-2 Fact Sheet, 2016. UNEP, WMO, UNCCD. “Global Assessment of Sand and Dust Storms.” United Nations Environment Programme, Nairobi, 2016. Wang, S. G., D. B. Yang, Y. S. Zhou, T. Ba, and C. L. Xin. “Analysis on the Formative Causes of Sand-Dust Storms in the Northwest China during 3–12 April 1994.” Journal of Desert Research 15, no. 4 (1995): 332–338. Wang, X. M., C. X. Zhang, E. Hasi, and Z. B. Dong. “Has the Three Norths Forest Shelter- belt Program Solved the Desertification and Dust Storm Problems in Arid and Semiarid China?” Journal of Arid Environments 74, no. 1 (2010): 13–22. Wiggs, Giles F. S., Sarah L. O’Hara, Johannah Wegerdt, Joost Van Der Meer, Ian Small, and Richard Hubbard. “The Dynamics and Characteristics of Aeolian Dust in Dryland Central Asia: Possible Impacts on Human Exposure and Respiratory Health in the Aral Sea Basin.” Geographical Journal 169, no. 2 (2003): 142–157. Wilderson, Walter D. “Dust and Sand Forecasting in Iraq and Adjoining Countries.” Air Weather Service Scott AFB IL (1991). WMO. “Organizations Delivering SDS Forecasts. World Meteorology Organization” (2011). http://www.wmo.int/pages/prog/arep/wwrp/new/documents/Organizations_­ delivering_SDS_forecasts.pdf World Bank and Institute for Health Metrics and Evaluation (IHME). “The Cost of Air Pol- lution: Strengthening the Economic Case for Action.” Washington: World Bank Group (2016). Sources, Costs, and Solutions 33 World Health Organization. “Environmental Burden of Disease: Country Profiles” (2018). https://www.who.int/quantifying_ehimpacts/national/countryprofile/intro/en/ ———. “Meningococcal Meningitis” (2015). http://www.who.int/ mediacentre/factsheets/ fs141/en/ Yoo, Young, Ji Tae Choung, Jinho Yu, Do Kyun Kim, and Young Yull Koh. “Acute Effects of Asian Dust Events on Respiratory Symptoms and Peak Expiratory Flow in Children with Mild Asthma.” Journal of Korean Medical Science 23, no. 1 (2008): 66–71. Young, Anthony. “Agroforestry for Soil Conservation” (1989). Zhang, Xuelei, Lijing Zhao, Daniel Tong, Guangjian Wu, Mo Dan, and Bo Teng. “A Sys- tematic Review of Global Desert Dust and Associated Human Health Effects.” Atmo- sphere 7, no. 12 (2016): 158. Zoljoodi, Mojtaba, Ali Didevarasl, and Abbas Ranjbar Saadatabadi. “Dust Events in the Western Parts of Iran and the Relationship with Drought Expansion over the Dust- Source Areas in Iraq and Syria.” Atmospheric and Climate Sciences 3, no. 03 (2013): 321–336. 34 Sand and Dust Storms in the Middle East and North Africa (MENA) Region GLOSSARY Aerosol  A collection of airborne solid or liquid particles, with a typical size between 0.01 and 10 micrometres (μm), that resides in the atmosphere for at least several hours. Aerosols may be of either natural or anthropogenic origin. Aerosol index  An indicator that detects the presence of UV-absorbing aerosols such as dust and soot. Agroforestry  The intentional integration of trees and shrubs into crop and animal farm- ing systems to create environmental, economic, and social benefits. Albedo  The proportion of the incident light or radiation that is reflected by a surface. Alkalinity  The accumulation of sodium ions on exchange surfaces of soils, resulting in high pH values and often collapse of soil structure due to dispersion of clays. Allergic rhinitis  Hay fever, a type of inflammation in the nose which occurs when the immune system overreacts to allergens in the air. Anthropogenic  As a result of human activity. Anticyclone  A weather system with high atmospheric pressure at its center, around which air slowly circulates in a clockwise (northern hemisphere) or counter-clockwise (southern hemisphere) direction. Anticyclones are associated with calm, fine weather. Aspergillosis  A condition in which certain fungi infect the tissues, especially the lungs. Biogeochemical cycles  The fluxes of chemical elements among different parts of the earth: from living to non-living, from atmosphere to land to water, and from soils to plants. Cerebrovascular diseases  A group of conditions that affect the circulation of blood to the brain, causing limited or no blood flow to affected areas of the brain. Chronic Obstructive Pulmonary Disease  The name for a collection of lung diseases including chronic bronchitis, emphysema, and chronic obstructive airways disease. Clay  Soil or sediment particles of diameter less than 2 microns. Cyclone  A system of winds rotating inward to an area of low atmospheric pressure, with a counter-clockwise (northern hemisphere) or clockwise (southern hemisphere) circulation; cyclones are associated with tropical storms. Sources, Costs, and Solutions 35 Deposition  Settling of dust onto the land surface through natural settling of particles from the atmosphere (dry) or in precipitation (wet). Desertification  When individual land degradation processes, acting locally, combine to affect large areas of drylands (UNEP 2007). Dust bowl  An area of land where vegetation has been lost and soil reduced to dust and eroded, especially as a consequence of drought or unsuitable farming practice. Dust haze  Dust which resides in the atmosphere from a previous dust storm. Dust storm  The result of terminal winds raising large quantities of dust into the air and reducing visibility at eye level (1.8 meters) to less than 1,000 meters. Emphysema  A condition in which the air sacs of the lungs are damaged and enlarged, causing breathlessness. Entrainment  The process of particle lifting by the agent of wind erosion. Ephemeral water body  A water body that dries up periodically. Erodibility  The inherent yielding or non-resistance of soils and sediments to wind erosion. Erosivity  A measure of the capacity of wind to cause soil or sediment erosion. Fetch  The length of unobstructed terrain over which the wind flows. Hydrologic  Associated with water bodies that dry out during some periods, including shorelines, river beds, ephemeral water bodies, and inland water features. Ischemic heart disease  Also known as coronary artery disease, a group of diseases that includes: stable angina, unstable angina, myocardial infarction, and sudden cardiac death. Land Degradation Neutrality  A policy supported by the UNCCD to maintain or improve the amount of healthy and productive land resources over time and in line with national sustainable development. It is incorporated into the Sustainable Development Goal target 15.3. Lidar  A detection system that works on the principle of radar but uses light from a laser. Loess  Sediment formed by the accumulation of wind-blown silt. Meningococcal meningitis  A bacterial infection that results in swelling and irritation (inflammation) of the membranes covering the brain and spinal cord; also known as cere- brospinal meningitis. Micron  A unit of millionth of a meter, or micrometer (μm). Mineral dust  Atmospheric aerosols originated from the suspension of minerals consti- tuting the soil, being composed of various oxides and carbonates. Mitigation of climate change  Efforts to reduce or lessen climate change through, for example, reduction in greenhouse gas emissions. Mitigation of sand and dust storms  Efforts to reduce anthropogenic causes of SDS and to lessen the negative impacts of SDS on human well-being. 36 Sand and Dust Storms in the Middle East and North Africa (MENA) Region Monsoon  A seasonal prevailing wind in the region of South and Southeast Asia, blowing from the southwest between May and September and bringing rain (the wet monsoon), or from the northeast between October and April (the dry monsoon). Natural ecosystem  An ecosystem that occurs as it would without the influence of human beings. Natural ecosystems include deserts, grasslands, natural forests, lakes, and rivers. Normalized Vegetation Difference Index  A measure of green vegetation cover cal- culated from satellite image data. Phytoplankton  Microscopic marine plants. Phytoplankton provide the base of several aquatic food webs. Playa  Flat-bottomed depressions commonly found in interior desert basins and as “sabkhas” adjacent to coasts within arid and semiarid regions; in some locations, these are ­ periodically covered by water to form playa lakes, some of which are saline. Ephemeral, salt, or dry lakes are referred to as playas and playa lakes in North America; salinas, saladas, and salars in South America; chotts (shatts, shotts); sebkhas or sabkhas in the Middle East; boinkas in Australia; pans in southern Africa; or kavir, or gol in Asia. Primary productivity  The rate at which plants and other photosynthetic organisms produce organic compounds in an ecosystem. Radiative balance  The relationship between the amount of energy reaching the earth and the amount leaving it. Reduced or no tillage  A practice of minimizing soil disturbance and allowing crop res- idue or stubble to remain on the ground instead of being removed, burned, or incorporated into the soil. Reduced tillage practices may progress from reducing the number of tillage passes to stopping tillage completely (no or zero tillage). Risk factor  A factor that raises the probability of an adverse outcome. Salinization  Accumulation of water-soluble salts in soil. Sand  Soil or sediment particles of diameter greater than 63 microns. Sediment  Naturally occurring material that is broken down by processes of weathering and erosion, and is subsequently transported by the action of wind, water, or ice, and/or by the force of gravity acting on the particles. Shelter belt  Planting of one or more rows of trees or shrubs in such a manner as to pro- vide shelter from the wind and to protect soil from erosion. Silicosis  Lung fibrosis caused by the inhalation of dust containing silica. Silt  Soil or sediment particles of diameter between 2 and 63 microns. Soil  The unconsolidated mineral or organic matter on the surface of the earth that has been subjected to and shows effects of genetic and environmental factors of: climate (includ- ing water and temperature effects), and macro- and microorganisms, conditioned by relief, acting on parent material over a period of time. A product-soil differs from the material from which it is derived in many physical, chemical, biological, and morphological properties and characteristics. Sources, Costs, and Solutions 37 Soil conservation contours  The practice of tilling sloping land or leaving strips of land untilled along lines of consistent elevation in order to conserve rainwater and to reduce soil losses from surface erosion. Soil cover  The degree to which soil is covered and protected by vegetation, organic litter layers, or mulches. Source apportionment  The practice of deriving information about pollution sources and the amount they contribute to ambient air pollution levels. Surface roughness  Character of a surface that produces drag on wind; results in turbu- lent flow with efficient transfer of matter and energy. Sustainable land management  Practices and technologies that aim to integrate the management of land, water, biodiversity, and other environmental resources to meet human needs, while ensuring the long-term sustainability of ecosystem services and livelihoods. 38 Sand and Dust Storms in the Middle East and North Africa (MENA) Region 1818 H Street, NW Washington, D.C. 20433 USA Telephone: 202-473-1000 Internet: www.worldbank.org/environment