327
TA 6479-REG: Addressing Climate Change in the Asia and Pacific Region Building Climate Resilience in the Agriculture Sector - DRAFT – Not for Quotation International Food Policy Research Institute April 27, 2009

Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

TA 6479-REG: Addressing Climate Change in the Asia

and Pacific Region

Building Climate Resilience in the Agriculture Sector

- DRAFT – Not for Quotation

International Food Policy Research Institute April 27, 2009

Page 2: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

ii

TABLE OF CONTENTS

CHAPTER 1. INTRODUCTION AND OVERVIEW ................................................................6

Key Messages ..................................................................................................................................6

1.1 Introduction ............................................................................................................................6

1.2 Key indicators related to agriculture and climate in the Asia and Pacific region ............7

1.3 Asian DMCs under a changing climate ..............................................................................10 1.3.1 Temperature and Precipitation .....................................................................................10 1.3.2 Frequency and severity of climate extremes ................................................................11

1.4 Conceptual framework ........................................................................................................12 1.4.1 Exposure and sensitivity ..............................................................................................12 1.4.2 Adaptive capacity.........................................................................................................13

1.5 Moving from vulnerability to resilience .............................................................................14 1.5.1 Resiliency approach .....................................................................................................14 1.5.2 Strategies for building resilience .................................................................................15

1.6 Conclusion .............................................................................................................................17

CHAPTER 2. IMPACTS OF CLIMATE CHANGE ON FOOD SECURITY ......................19

Key Messages ................................................................................................................................19

2.1 Introduction ..........................................................................................................................20

2.2 Food availability ...................................................................................................................21 2.2.1 Impacts of climate change on agricultural potential and crop yields—literature

review ...........................................................................................................................21 2.2.2 Impacts of climate change on agriculture and crop yields—combined IFPRI

crop modeling, neural network and IMPACT results ..................................................29 2.2.3 Impacts of climate change on trade and GDP—literature review ..............................32 2.2.4 Impacts of climate change on food prices and net trade -- combined IFPRI crop

modeling, neural network and IMPACT results ..........................................................34

2.3 Food accessibility ..................................................................................................................35 2.3.1 Market dimension ........................................................................................................35 2.3.2 Nonmarket dimension ..................................................................................................37

2.4 Food utilization .....................................................................................................................42 2.4.1 Global burden of disease ..............................................................................................42 2.4.2 Overall health impacts .................................................................................................45 2.4.3 Health impacts on vulnerable people ...........................................................................45

Page 3: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

iii

2.4.4 Impacts of climate change on calorie availability and malnutrition -- combined IFPRI crop modeling, neural network and IMPACT results .......................................47

2.4.5 Role of health infrastructure ........................................................................................49

2.5 Food system stability ............................................................................................................50 2.5.1 Impacts of climate extreme events ...............................................................................50 2.5.2 Potential for conflicts ...................................................................................................52

2.6 Conclusion ............................................................................................................................53

CHAPTER 3. VULNERABILITY OF ASIA AND PACIFIC COUNTRIES TO CLIMATE CHANGE ..................................................................................................................55

Key Messages ................................................................................................................................55

3.1 Factors affecting vulnerability to climate change in the Asia and Pacific region ..........55

3.2 Results from vulnerability assessments for the Asia and Pacific region .........................56

3.3 Vulnerability indicator for Asia and the Pacific region ....................................................63

3.4 Conclusion .............................................................................................................................64

CHAPTER 4. OPPORTUNITIES FOR MITIGATION AND SYNERGIES WITH ADAPTATION AND SUSTAINABLE DEVELOPMENT .....................................................65

Key Messages ................................................................................................................................65

4.1 Introduction ..........................................................................................................................65

4.2 Global and regional emissions trends and sources ............................................................66 4.2.1 Agricultural soils ..........................................................................................................67 4.2.2 Livestock and manure management .............................................................................68 4.2.3 Rice cultivation ............................................................................................................68

4.3 Mitigation strategies in the agricultural sector .................................................................68 4.3.1. Carbon sequestration ....................................................................................................69 4.3.2 Bioenergy .....................................................................................................................70 4.3.3. On-farm mitigation ......................................................................................................71 4.3.4 Summary of technical mitigation potential ..................................................................72 4.3.5 Economic potential of mitigation options ....................................................................72 4.3.6 Summary of economic mitigation potential .................................................................74

4.4 Institutional barriers to mitigation in agriculture in the Asia and Pacific region .........74

4.5 Integrating mitigation and adaptation in sustainable development pathways ...............75 4.5.1 Synergies between mitigation and adaption ................................................................76

Page 4: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

iv

4.5.2 Suggestions for mainstreaming mitigation and adaptation in sustainable development pathways ...........................................................................................................77

4.6 Conclusions ...........................................................................................................................79

CHAPTER 5. POLICIES, INVESTMENTS, AND INSTITUTIONS FOR ENHANCED RESILIENCE OF DEVELOPING ECONOMIES IN ASIA AND PACIFIC IN THE FACE OF CLIMATE CHANGE ...............................................................80

Key messages ................................................................................................................................80

5.1 Introduction ..........................................................................................................................81

5.2 Agricultural adaptation for the Asia and Pacific region ..................................................83 5.2.1 Local coping strategies ................................................................................................85 5.2.2 Innovative adaptation to climate change ......................................................................88

5.3 Strengthening important ongoing development initiatives to support climate change adaptation and mitigation .....................................................................................100 5.3.1 Secure property rights ................................................................................................100 5.3.2 Agricultural policies ...................................................................................................101 5.3.3 Trade policies .............................................................................................................101 5.3.4 Other environmental policies .....................................................................................102 5.3.5 Social protection ........................................................................................................104 5.3.6 Financial markets: The role of microfinance .............................................................105 5.3.7 Disaster preparedness.................................................................................................105

5.4 Implementing Climate Change Adaptation Policies .......................................................106 5.4.1 Mainstreaming climate change and adaptation into development planning ..............106 5.4.2 National Adaptation Plans of Action (NAPA)...........................................................107 5.4.3 Significant new investments ......................................................................................108 5.4.4 Cost of adaptation for the Asia-Pacific region ...........................................................109 5.4.5 Financing adaptation ..................................................................................................110

5.5 Reforming climate-change related governance and institutions ....................................112 5.5.1 Civil society ...............................................................................................................114

5.7 Adaptation policy recommendations ................................................................................117

CHAPTER 6. CONCLUSIONS AND POLICY RECOMMENDATIONS..........................118

REFERENCE LIST ...................................................................................................................127

TABLES ......................................................................................................................................151

FIGURES ....................................................................................................................................189

Page 5: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

v

ANNEXES ..................................................................................................................................238

Annex 1.1. List of ADB’s Developing Member Countries, by Region (as classified in the Annual Asian Development Outlook Publication) ...........................................................238

Annex 2.1. Models Used for Agricultural Climate Change Impact Analysis. ......................243

Annex 2.2. IFPRI’s Climate Change Modeling Framework..................................................249

Annex 3.1 Parameters for agricultural employment as share of total employment (reflecting sensitivity to climate change) ..................................................................................257

Annex 3.2. Poverty incidence reflecting relative adaptive capacity in the Asia and Pacific region. .............................................................................................................................259

Annex 5.1. Local coping strategies as adaptation tools to mitigate the impacts of climate change in agriculture. ...................................................................................................260

APPENDIX .................................................................................................................................278

Appendix 1. Country case studies ............................................................................................278

Page 6: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

6

CHAPTER 1. INTRODUCTION AND OVERVIEW

Key Messages

Agriculture is important for all developing member countries (DMCs) of the Asian Development Bank: More than 60 percent of the economically active population and their dependents—2.2 billion people—rely on agriculture for their livelihood.

While agriculture is crucial for the region’s food security and the backbone for much of the employment in the region, farming systems vary significantly across region, ranging from the relatively dry wheat-producing areas of Central Asia to the very wet rice producing lands of Southeast Asia. Similarly, support afforded to agriculture and agricultural technologies employed vary significantly across the various countries. The heterogeneity in farming systems will require targeted interventions for adaptation (and mitigation) of climate change.

Given that the Asia and Pacific region accounted for 43 percent of global crop production in 2000 and is expected to account for one third of total cereal demand and two thirds of total meat demand over the next several decades, and furthermore accounts for significant net cereal exports, particularly rice, climate change impacts on this region will impact food security not only regionally, but also globally.

All regions in the Asia-Pacific region are expected to get warmer.

The region is expected to experience an increase in frequency of climate extreme events.

The Asia and Pacific region is expected to generally get wetter, with the exception of Central Asia. However, rainfall tends to be heavier during wet periods, increasing the risk of floods, while dry seasons will remain dry or get drier.

Moreover, the Asia and Pacific region is considered particularly vulnerable to sea level water rise and glacier melt.

Developing resilience to climate change in the Asia Pacific region and elsewhere requires working on all three fronts affecting vulnerability: exposure, sensitivity and adaptive capacity.

Strategies to build resilience including: Minimize reactive coping and promote proactive adaptation, plan for uncertainty, build resilience for agroecological systems, strengthen local adaptive capacity, support sustainable development, and support measures that promote efficiency, productivity, and alternatives to agriculture will be crucial for the region to adapt to climate change.

1.1 Introduction

Climate change is threatening food production systems and therefore the livelihoods and food security of billions of people who depend on agriculture in the Asia and Pacific region. Evidence shows that marginalized populations will disproportionately suffer the impacts of climate change in comparison with wealthier, industrial countries (IPCC 2007a). Relatively poorer countries will not only experience more severe impacts, but these countries often also lack the resources to prepare for and cope with environmental risks. Agriculture is the sector most vulnerable to

Page 7: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

7

climate change because of its high climate dependence and because people involved in this sector depend to be poorer compared to their urban compatriots. Among the developing member countries (DMCs) of the Asian Development Bank more than 60 percent of the economically active population and their dependents—2.2 billion people—rely on agriculture for their livelihoods (FAOSTAT).1 A list of DMCs is presented in Annex 1.1.

Climate change is already evident in a number of ways. Consistent warming trends and more frequent and intense extreme events have been observed across Asia and the Pacific in recent decades. In line with these trends, climate change scenarios consistently project temperature increases across Asia and the Pacific. There is much less certainty and agreement among models on rainfall variability and frequency and intensity of extreme climate events, but extreme events such as cyclones, floods, hailstorms, and droughts are generally expected to increase in frequency and severity across the region, as well as in specific areas.

The observed and projected future effects are diverse and geographically differentiated, creating uncertainty that makes the task of preparing for climate change impacts difficult. Just as the impacts will be varied, so will each community’s ability to respond to changes in environmental conditions. Expected climate effects on agroecologies will consist of both rapid and catastrophic shifts that cause crop failure and immediate food shortages and longer-term shifts such as slow changes in mean temperature and increased inter-annual and seasonal climate variability. Dealing with the short- and longer-term impacts on agricultural systems will require improving the understanding of vulnerable production systems and building the capacity to adapt to these changes.

Climate change will place an additional burden on efforts to meet long term development goals in the Asia and Pacific region. Slow agricultural productivity growth, slow income growth, and problems maintaining food security already pose challenges to many countries in the region. Therefore, meeting any development goal requires increasing investments due to climate change. The challenge of reducing the risks of climate change for agriculture is thus immense. The overarching goal of this report is to provide a framework for approaching this challenge by establishing baselines of knowledge of climate impacts and plausible theories about how to build longer-term adaptive capacity and resilience. The book will both summarize results from the literature and make use of novel quantitative assessment tools. The remaining sections of this chapter review key indicators related to agriculture and climate change in the Asia and Pacific region and introduce a conceptual framework for moving from vulnerability to resilience.

1.2 Key indicators related to agriculture and climate in the Asia and Pacific region

Approximately 55 percent of the world’s population resides in ADB’s developing member countries (for a listing of countries by region, see Annex 1.1).2 Agriculture—the principal source of livelihood for more than 60 percent of the population of the Asia and Pacific region—is extremely vulnerable to climate change. Therefore, the effects of climate change on food production systems will directly affect the primary source of income for billions of people in the region. Moreover, perturbations in food supply will have implications for the wider population

1 Agriculture, as defined by the Food and Agriculture Organization of the United Nations (FAO), includes farming, fishing,

hunting, and forestry. 2 For the remainder of this section, statistics for this region follow ADB member-country classification and are from

FAOSTAT (2009).

Page 8: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

8

who are net food purchasers, as well as raising the importance of trade both within the region and around the world. Given that the Asia and Pacific region accounted for 43 percent of global crop production in 2000 and is expected to account for one third of total cereal demand and two thirds of total meat demand over the next several decades, and furthermore accounts for significant net cereal exports, climate change impacts on this region will impact food security globally.

Tables 1.1 and Table 1.2 present important indicators that can be used to assess the agriculture sector’s vulnerability to climate change in Asia and Pacific countries. Table 1.1 presents historic annual climate means as well as indicators of agricultural dependence and poverty in the Asia and Pacific region. Table 1.2 presents the main crop and livestock products, by tonnage, produced in each of the ADB DMCs for the year 2007. These production statistics are useful in understanding differences in main crop and livestock products among regions, and help to give an indication of the type of production technologies that maybe impacted by climate change. For example, China accounts for almost one fifth of global maize and wheat production and for 29 percent of global rice production (2007 values). Impacts on the country’s production system from climate change therefore require close monitoring. Knowledge on Asian farming systems is also important for mitigation and adaptation in terms of food security and livelihoods. An overview by region is presented below.

Central Asia

The Central Asia region is the second least populated region in Asia and the Pacific, with 75 million inhabitants. Among the countries in Central Asia, Uzbekistan has the largest population (36 percent of the total). More than half of the region’s population lives in rural areas (42 million), but less than a quarter of the people (17 million) derive their livelihoods from agriculture. Population density on arable land is moderate, with an average of 245 people per square kilometer (km2). Despite being relatively land abundant, the importance of agriculture to GDP has been declining, with the exception of Turkmenistan. The Republic of Georgia has made the most significant strides in this category, reducing the importance of agriculture to GDP from 52 percent in 1995 to 13 percent in 2006. In Tajikistan and Turkmenistan, where agriculture remains a significant share of GDP, employment in agriculture remains high—around 30 percent of the economically active population. Finally, the proportion of undernourished in the total population has been falling since 1995, with the exception of Uzbekistan, which has nearly tripled the percentage from 5 to 14 percent.

Rainfall in the region is lowest compared to the other regions, averaging less than 500 mm annually. As a result, more than half of the countries in this region irrigate at least 50 percent of cropland, which is crucial for the region’s food production and employment. Key crops are wheat, which is chiefly produced in Kazakhstan, potatoes, and seed cotton, while main livestock products are milk and beef.

East Asia

The East Asia region is the second largest region in terms of population, with 1.4 billion inhabitants, most of whom reside in China. Nearly 60 percent of the people (792 million) live in rural areas and about the same proportion (847 million) is sustained by some form of agriculture. Rural population density on arable land is high—at 559 persons per km2 in China—but below some of the land-scarcer countries in South and Southeast Asia. The importance of agriculture in

Page 9: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

9

GDP has been declining across the region. While agriculture accounts for only 12 percent of GDP in China, nearly 64 percent of the economically active population is employed in agriculture. Finally, while food security has been improving in the region overall, nearly 30 percent of the population of Mongolia is undernourished. Given significant land scarcity in the region, several of the countries in this region, including China, Japan, and South Korea, have started to purchase or lease land for food production in other parts of Asia (Indonesia), in Africa, Eastern Europe, and in Latin America.

Rice is the major crop in the region, with China producing 187 million metric tons in 2007. Other key crops include maize, pig meat and milk in China, and rice, fresh vegetables, and milk in Korea. Irrigated land has a moderate presence in China and Korea, as cereals other than rice are often not irrigated. Rainfall is lowest in Mongolia and moderate in China—with large variations within the country—while Korea receives the most with an average of over 1,300 mm per year.

South Asia

The South Asia region comprises eight countries, the largest of which is India. More than one-quarter of the population of the developing world is found in South Asia. Of this population of more than 1.6 billion people, more than two-thirds (70 percent) live in rural areas. Approximately 787 million people can be classified as agriculture dependent. Given the high population density in this region, there is only about 0.16 hectare of agricultural land per capita. Rural population density per km2 of arable land is highest in Bangladesh and Sri Lanka, each with over 1,000 inhabitants, with India, Nepal, and Pakistan having relatively moderate density. The importance of agriculture to GDP remains high in the region, only declining slightly from 1995 to 2006. As a result, employment in agriculture is also high, with close to 50 percent or more of the population dedicated to this sector (with the exception of the Maldives). Finally, the proportion of undernourished in the population averages over 20 percent, making South Asia the least food secure region within the Asia and Pacific and the world.

Average rainfall varies across South Asia, with Bangladesh receiving the largest amount and Afghanistan receiving the least and a regional average of about 1,300 mm per year, which is on par with averages for India and Nepal. As a result, irrigation coverage is high in the region, varying from over 80 percent of cropland in Pakistan to at least 30 percent of cropland in Afghanistan, Sri Lanka and India. Irrigation supports the production of major crops such as sugar cane, rice, and wheat in India and Pakistan, and rice in Nepal.

Southeast Asia

The Southeast Asia region comprises nine countries and 564 million people, with more than 40 percent of the population (229 million) living in Indonesia. More than half the region’s population resides in rural areas. Approximately 46 percent of people (257 million) rely on agriculture for their livelihoods. Rural population density per km2 of arable land is moderate, ranging from 588 inhabitants in Indonesia to less than 300 in Cambodia. The importance of agriculture to GDP has been declining in the region; however it still contributes 30 percent in Cambodia and over 40 percent in Lao PDR. Finally, undernourishment in the region has been declining since 1995, but still averages 18 percent of the population, with 26 percent of the population of Cambodia classified as malnourished.

Page 10: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

10

Southeast Asia receives over 2,000 mm of rainfall, on average, each year, which is second only to Pacific region. As a result, agricultural area remains largely rainfed, whereas irrigated cropland averages approximately 17 percent of the total. These conditions favor crops such as rice, sugar cane and oil palm fruit, which are the dominant crops in this region in terms of tonnage of production. In terms of livestock production, poultry and chicken are of greatest importance for Indonesia, while the Philippines produces a significant amount of pig meat.

The Pacific

The smallest region in terms of population is the Pacific, with 9.4 million inhabitants. Eleven of the 14 countries in this region have less than 500,000 inhabitants. The most populous country in the region is Papua New Guinea, with 5.9 million people. More than 80 percent of the population of these islands can be classified as rural, and about 67 percent (6.1 million) are dependent upon agriculture for their livelihoods. Data for this region on irrigated cropland, undernourishment, and the importance of agriculture in GDP are scarce. Data from Papua New Guinea indicates, however, that the percentage of agriculture in GDP has been rising, from 32 percent in 1995 to 42 percent of GDP in 2005. In addition, the proportion of the population employed in agriculture averages close to 40 percent. Finally, calorie availability has improved slightly since 1995, rising from 2,560 to 2,660 kilocalories per person per day in 2005.

1.3 Asian DMCs under a changing climate

1.3.1 Temperature and Precipitation

The IPCC’s non-mitigation scenarios indicate that the equilibrium global mean SAT (surface air temperature) warming for a doubling of atmospheric carbon dioxide (CO2) is likely to range from 2oC to 4.5oC (degrees Celsius) (Meehl et al. 2007). The HadCM3/A2a scenario results used for the quantitative analyses in this volume also suggest that all countries in Asia and the Pacific are expected to warm during the 21st century (Table 1.3; Figure 1.1). Temperature increases are expected to be largest in Central Asia, averaging 3.5 (oC) warmer in 2050 relative to their historical mean. East, South, and Southeast Asia are expected to warm on average over 2oC by 2050, with the more northern regions experiencing the greatest increases. The Pacific region is predicted to experience the least warming given their proximity to the ocean, with warming averaging 1.3oC, by 2050. These predictions are similar to Christensen et al. (2007) who found that warming is expected to be similar to the global mean in Southeast Asia (mean warming between 1980-1999 and 2080-2099 of 2.5oC); well above the global mean in Central Asia, the Tibetan Plateau, and Northern Asia; above the mean in East and South Asia; and below the mean in the Pacific Islands (Christensen et al. 2007). It should be noted that Christensen et al. (2007) refer to various scenarios and models, while we focus on one scenario and that the time periods are slightly different.

IPCC’s projections also indicate that it is very likely that there will be longer and more intense summer heat waves/hot spells in East Asia, and there will be fewer cold days in East Asia and South Asia (Christensen et al. 2007). Trends in that direction can already be perceived in parts of Asia. In Southeast Asia and the Pacific Region, analyses of daily temperature in the period from 1961 to 1998 for 15 countries (91 stations) indicates significant increases in the annual number of hot days and warm nights and significant decreases in the annual number of

Page 11: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

11

cool days and cold nights. These trends were considerably consistent across the region (Manton et al. 2001).

While there is some confidence regarding temperature changes in the tropics as a result of climate change (at least regarding direction), there is far greater uncertainty about precipitation changes (Christensen et al. 2007). For many sub-regions, there is even a lack of consistency in predicting the overall trends for precipitation. The inclusion of precipitation predictions in climate scenarios is, however, extremely important. Studies show that much of the direction and magnitude of impacts of climate change on agriculture and livestock are conditioned to assumptions regarding precipitation changes in conjunction with temperature increases (Kurukulasuriya and Ajwad 2007; Batima 2006; Fischer et al. 2005). Therefore, it is important to interpret results with an understanding of the underlying predictions in precipitation and temperature changes.

In the past decades, rainfall trends in the Asia and the Pacific region have varied according to sub-region. A review of studies about observed past and present climate trends and variability indicate decreasing trends in annual mean rainfall in North-East and North China, the coastal belts and arid plains of Pakistan, parts of North-East India, the east coast of India, Indonesia, and the Philippines (Cruz et al. 2007; Preston et al. 2006). Increasing trends of annual mean rainfall have been observed in western China, the Changjiang (River Yangtze) Basin, the southeastern coast of China, the Arabian Peninsula, Bangladesh, and along the western coasts of the Philippines (Bates et al. 2008; Preston et al. 2006).

The IPCC Fourth Assessment’s models project an increase in annual precipitation in most parts of Asia during this century, with larger and more consistent increases in North and East Asia. Projections suggest that boreal winter precipitation is very likely to increase in northern Asia and the Tibetan Plateau and likely to increase in eastern Asia and the southern parts of Southeast Asia. An exception is Central Asia, where a decrease in precipitation is expected in the summer. In that region, the projected decrease in mean precipitation is expected to cause an increase in the frequency of dry spring, summer, and autumn seasons (Christensen et al. 2007). Increases in precipitation levels for most Asian countries and decreases in Central Asian countries are confirmed by a recent study that makes projections at the country level (Cline 2007). However, increases in annual rainfall do not necessarily mean that regions will have fewer drought events, as in many cases rainfall tends to be heavier during wet periods, increasing the risk of floods, while dry seasons continue to not get enough rain.

Table 1.3 presents changes in precipitation for the Hadley Model, A2a scenario, compared to the historical mean, while Figure 1.2 depicts these changes graphically. Rainfall is expected to increase slightly in all regions with the exception of parts of Central Asia, Cambodia and Malaysia. East and South Asia and the Pacific will experience increased rainfall of approximately 10 percent above the historical mean in 2050, while rainfall in Southeast Asia will only increase slightly. These projections are consistent with trends predicted by Cline (2007b).

1.3.2 Frequency and severity of climate extremes

Table 1.4 indicates country vulnerability to sea-level rise, floods, droughts, and storms based on disaster frequency from 1900 to 2008. All countries in the Asia and Pacific region have experienced weather related disasters in the past hundred years. In addition, extreme climate events such as floods, droughts, and typhoons have increased in both frequency and severity in

Page 12: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

12

many regions of the world (Sanker, Nakano, and Shiomi 2007; IPCC 2007a; Cruz et al. 2007). Along with environmental degradation, land use changes, and high population density, climate change is considered one of the main causes of these changes. For instance, a study shows that droughts in Southwest Asia, southern Europe, and the United States in 1998-2002 were linked to cold sea surface temperatures (SSTs) in the eastern tropical Pacific and unprecedented warm SSTs in the western tropical Pacific and Indian oceans. Climate models indicate that each of those regions contributed to generate a synchronized drought. Despite the fact that ENSO3 is a natural phenomenon that has occurred for a long time, the warming of the Indian Ocean and the western Pacific Ocean was beyond that expected from natural variability and partly due to the ocean’s response to increased greenhouse gases (Hoerling and Kumar 2003). Therefore, parts of Asia have experienced longer heat waves and more frequent and intense droughts. In Southeast Asia, extreme weather events associated with El Niño have also increased in frequency and intensity in recent decades (Cruz et al. 2007). Damage caused by cyclones has also significantly risen in countries such as India, China, the Philippines, Japan, Viet Nam, Cambodia, Iran, and the Tibetan Plateau (Cruz et al. 2007).

Several models predict an increase in the intensity of heavy rainfall and winds in South Asia (over the Arabian Sea and the tropical Indian Ocean; northern Pakistan; and northwest, northeast, west coast, and west central India; Bangladesh; and Myanmar), East Asia (Japan, South Korea, China) and Southeast Asia (Christensen et al. 2007). For the Pacific Islands, by 2030 and 2070, models consistently predict more intense cyclones (increase in wind speed) and increases in rainfall greater than 10 percent in the Islands east of Papua New Guinea (Solomon Islands, Kiribati, and Tuvalu) (Preston et al. 2006;World Bank 2000). In particular, Pacific Atoll countries are likely to see more intense rainfall events and droughts. In 2080, flood risk is expected to be 200 times greater than at present for these countries (Barnett and Adger 2003). Additional details on climate extreme events and the linkage to vulnerability are presented in Chapter 3.

1.4 Conceptual framework

Figure 1.3 presents the conceptual framework for building climate change resilience in human systems. A community’s ability to become resilient to climate change is determined by the nature of the impacts and the community’s capacity to adapt. The magnitude of impacts is influenced by exposure and sensitivity to climatic variability and change. In addition, varied factors determine adaptive capacity, from social networks to the level of access to economic resources. Therefore, building adaptive capacity through increasing access to knowledge and resources and reducing the severity of impacts through emissions abatement are key entry points for reducing vulnerability and building resilience to climate change.

1.4.1 Exposure and sensitivity

Exposure and sensitivity have been used in the literature to characterize the biophysical impacts of climate change on agroecological systems (Tubiello and Rosenzweig 2008; Moss et al. 2001). Exposure encompasses the spatial and temporal dimensions of climate variability, such as

3 ENSO - El Niño Southern Oscillation is a disruption of the ocean-atmosphere system that occurred in tropical Pacific but

with significant impacts on global weather such as redistribution of rains with extreme flooding and droughts (Neelin et al. 1998).

Page 13: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

13

droughts, and sensitivity refers to the resiliency of the agroecological system to withstand the impacts, without conscience efforts by manages at adaptation. Levels of exposure and sensitivity to climate change will affect crop yields, water availability, pest populations, and crop calendars. Chapter 2 includes a review of the results of existing agroclimatic models that seek to capture exposure and sensitivity to climate change and determine the levels of impacts on key biophysical variables, such as yields and water availability, and how they affect food security. This assessment will provide a good understanding of the impacts of climate change and variability on food security and agriculture, which is essential to improve adaptation and mitigation strategies and so to reduce vulnerabilities.

Vulnerability to climate change depends not only on exposure to climate events but also on physical, environmental, socioeconomic, and political factors that influence how sensitive countries will be to a changing climate and how they will be able to cope. Each of the three components defining vulnerability to climate change—exposure, sensitivity, and adaptive capacity—requires several strategies in order to reduce vulnerabilities of agriculture and rural communities in the countries of the Asia and Pacific region. Mitigation and adaptation measures are essential as a way to reduce the extent of global warming, to reduce sensitivity of countries, and to improve countries’ capacity to adapt to a changing climate.

Chapter 3 reviews the vulnerability of countries in the Asia and Pacific region based on composite indicators reflecting exposure, sensitivity, and adaptive capacity. In addition, the chapter provides a brief review of vulnerability assessments to climate variability and climate change, sea-level rise, and land and water degradation. Based on these assessments, we put together a simple but consistent vulnerability indicator combining elements of exposure to climate change, sensitivity to climate change, and adaptive capacity.

1.4.2 Adaptive capacity

The third dimension of vulnerability weighs the biophysical impacts against the capacity of a society or human system to manage those impacts. This aspect of vulnerability is most difficult to conceptualize, because many socioeconomic variables determine adaptive capacity. Institutional factors such as property rights and political stability will influence the extent to which farmers and other stakeholders can mobilize and gain access to pooled resources and knowledge. For example, government-provided extension services will influence a farmer’s knowledge of alternative technologies, and property rights provide an incentive for continued investment. Economic aspects will shape the level of investment and planning, as well as how much access to inputs such as fertilizer and irrigation a farmer may have.

Achieving enhanced resilience in the face of climate change will require enhancing the adaptive capacity of countries in Asia and the Pacific as well as implementing appropriate adaptation investments, policies, and institutions. Moreover, mitigation measures can support adaptation options and provide much-needed funds for further adaptation (Bryan et al. 2008; FAO 2009). Opportunities for integrating mitigation and adaptation into sustainable development pathways are discussed in Chapter 4.

Adaptation measures should be targeted to countries, sectors, and people most vulnerable to the adverse impacts of climate change—that is, those most exposed to and most sensitive to the adverse impacts of climate change and those with the least adaptive capacity (Figure 1.3).

Page 14: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

14

Chapter 5 develops a framework for prioritizing adaptive measures and building adaptive capacity.

1.5 Moving from vulnerability to resilience

Resilience is used to describe the magnitude of a disturbance that a system can withstand without crossing a threshold into a new structure or dynamic. A number of factors contribute to a system’s ability to become resilient, including economic and natural resources, knowledge, and the level of sophistication of institutional processes, all of which broadly describe a system’s adaptive capacity. Building resiliency to climate change requires simultaneously building resilience in human systems and in the interlinked ecosystems upon which they depend. Key concepts describing climate change resilience are presented in Box 1.1.

1.5.1 Resiliency approach

Humans rely on ecosystem services for survival, yet the institutional ability to manage ecosystems sustainably has not kept pace with the changes taking place within these systems. Socioeconomic institutions have considered ecosystems and the services they provide to be infinite and largely in a steady cycle of regeneration. This attitude has led to the creation of economic instruments and incentives that use ecosystems deterministically, from extraction to consumption. The concept of resiliency, however, challenges the idea of linear planning and recognizes that social and environmental systems are interlinked, complex adaptive systems that are process dependent rather than input dependent and self-organizing rather than predictable (SRI 2009). The lens of resiliency is useful for analyzing climate change in that it recognizes the human-ecological system as complex, unpredictable, and dynamic and tailors institutional measures and responses based on this principle. This approach makes clear that the uncertainty surrounding climate change is the most important issue. As a result, planning for uncertainty will be a major theme in this book. Although some changes in ecosystems—such as long-term temperature increases—may be smooth and linear, short-term climatic variability is erratic and often severe. Moreover, the spatial distribution of climate impacts will vary and are uncertain.

Box 1.1. Key concepts in building climate change resilience

Adaptation—an adjustment made in response to a perceived change in a human or natural system in order to reduce vulnerability, build resilience, or both. Adaptation can be proactive (anticipatory) or reactive, and planned or autonomous.

Adaptive capacity—the ability of institutions and individuals to avoid potential damage, to take advantage of opportunities, or to cope with consequences of change.

Resilience—describes the magnitude of a disturbance that a system can withstand without crossing a threshold into a new structure or dynamic.

Ecosystem resilience—“a measure of how much disturbance (like storms, fire, or pollutants) an ecosystem can handle without shifting into a qualitatively different state. It is the capacity of a system to both withstand shocks and to rebuild itself if damaged” (SRI 2009).

Social resilience—“the ability of human communities to withstand and recover from stresses, such as environmental change or social, economic, or political upheaval” (SRI 2009). This idea is similar to adaptive capacity.

Sensitivity—“the degree to which a system is affected, either adversely or beneficially, by climate

Page 15: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

15

variability or change. The effect may be direct (e.g., a change in crop yield in response to a change in the mean, range, or variability of temperature) or indirect (e.g., damages caused by an increase in the frequency of coastal flooding due to sea-level rise)” (IPCC 2007a).

Sustainable development—“The goal of sustainable development is to create and maintain prosperous social, economic, and ecological systems. Sustainable development has also been described as fostering adaptive capabilities and creating opportunities. This definition comes from combining sustainability—the capacity to create, test, and maintain adaptive capability—and development—the process of creating, testing, and maintaining opportunity” (Holling 2001 as quoted in RA 2009).

Synergy—“When the combined effect of several forces operating is greater than the sum of the separate effects of the forces” (MA 2005).

Vulnerability—“the degree to which a system is susceptible to, or unable to cope with, adverse effects of climate change, including climate variability and extremes. Vulnerability is a function of the character, magnitude, and rate of climate variation to which a system is exposed, its sensitivity, and its adaptive capacity” (IPCC 2007a). “Vulnerability is often denoted the antonym of resilience” (SRI 2009).

Enacting proactive adaptation measures to build resilience in food production, food security, and the livelihoods of the poor requires an understanding of the strategies that can reduce the vulnerability of the agricultural sector to the effects of climate change and that can build resilience among farmers and those who depend directly on the sector for their livelihoods. For example, key crops and production systems will need to be identified that are important both in maintaining food security and in generating income, as well as those production systems that will respond ideally to defensive investments. The determination of the best-bet investments depends on many factors and is more an art than a science at this point. Given that the location and severity of climatic events are difficult to predict, it will be important to identify strategies that are robust in the face of uncertainty.

1.5.2 Strategies for building resilience

What types of policy measures are needed to increase resiliency given the uncertain future we face? In the short and long run, which kinds of institutions, actions, and processes build capacity to adapt? These policy measures are discussed briefly here.

Minimize reactive coping and promote proactive adaptation. In the longer run, a goal of policy should be to shift toward proactive adaptation rather than merely reactive coping. Coping is typically defined as responding to short-term events such as weather disasters, whereas adaption responds to longer-term threats such as climate change. Both will be important strategies for farmers and the poor. Yet it is likely that short-term coping strategies in response to climate variability may have trade-offs with other socio-ecological goals. For example, expanding production into forested areas will decrease carbon sequestration and habitat.

Proactive adaptation seeks to minimize trade-offs between short-term coping and longer-term goals of sustainability while also building resilience to climate change in the long run. Proactive adaptation includes a range of measures, from adopting resource-conserving technologies and raising productivity to promoting alternative livelihoods. Farm-level and regional strategies include supporting information and technology sharing for decision support systems in order to raise competency and build decision-making capacity.

Page 16: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

16

Plan for uncertainty. Planning for uncertainty puts the focus on the process in which decisions are made instead of on projects that require large investments in infrastructure that are immutable or high in sunk costs. Supporting knowledge, coordination, collaboration, information exchange, and institutional responsiveness will be the backbone for building the broad set of technical skills needed to prepare, plan, and respond to a wide range of unpredictable contingencies. To that end, it will be important to investigate the types of policies that enhance social learning and build institutions’ adaptive capacity to deal with uncertainties in their local settings. This approach will require, in part, a detailed institutional assessment that highlights key areas of need and opportunity for building resiliency in institutions that support agricultural decision makers and, more generally, vulnerable populations.

Adaptive and flexible management will be essential, including the capacity to monitor the results of managers’ decisions and to subsequently modify actions as needed. The broadening nature and increasing severity of potential climate impacts in a given area and the unavoidable uncertainties associated with predicting these impacts requires innovative approaches to management and development that go beyond centralized prediction and control practices (Nelson et al. 2008; Pahl-Wostl 2007a). One approach—adaptive management, or adaptive governance—has received attention because it enables decision makers and resource managers to work with the inherent uncertainty associated with climate change (Pahl-Wostl 2007b; Brunner et al. 2005; Tompkins and Adger 2004; Folke et al. 2002).

Build resilience of agroecological systems. Building resilience in agroecological systems is critical because these systems determine our capacity to produce food and clean water. Whereas climate change will disturb the functioning of these systems in ways that could lead to severe losses of ecosystem functioning, such as desertification and soil degradation, building ecosystem resilience will enhance the capacity of these systems to withstand shocks and rebuild after damage. The adoption of resource-conserving technologies, such as rainwater harvesting, conservation tillage, and integrated crop, water, and pest management, will form the backbone of actions to sustain and enhance agroecological systems. In addition, policy measures that promote research and adoption of new drought- and heat-resistant crop varieties, strengthen water use productivity and performance, and promote synergies between adaptation and mitigation, will be most effective. These types of measures are discussed in Chapters 4 and 5.

Strengthen local adaptive capacity. Innovative responses to climate change, which will be key to agricultural adaptation, are already in development but have not been implemented on a wide scale. These responses include changes in agricultural practices for crop and livestock systems implemented as a result of decisions made by millions of farmers. Enhancing farmers’ ability to respond to climate variability and climate change will require significant improvements in developing and disseminating agricultural technologies targeted to the major evolving biotic and abiotic stresses generated by climate change. Improved crop varieties have the potential to be more drought-tolerant, to increase nutrient-use and water-use efficiency, and to decrease pesticide use. But new technologies, by themselves, are not sufficient to successfully address the challenges climate change poses for agriculture, including increased risks to production and household income. Such measures are discussed in more detail in Chapter 5.

To protect against devastating outcomes from agricultural failures due to weather and climate, programs and policies should be implemented to improve risk management and promote crop insurance, including climate-based insurance. These programs can also reduce risk aversion in farmers’ production decisions and thus enhance the potential for adoption of adaptive farming

Page 17: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

17

systems. A stable and supportive policy environment that makes those programs available and profitable is also a critical factor. Such a policy environment requires strengthening important ongoing development initiatives to support climate change adaptation, which have been implemented to varying degrees throughout the developing world. These initiatives include secure property rights; improved economic incentives and green markets; improved information collection, use, and dissemination; extension services; and enhanced social protection and fiscal resiliency.

Finally, effective implementation of this aggressive climate change adaptation agenda will require mainstreaming climate change and adaptation into development planning, reforming climate-related governance and institutions, and undertaking massive new investments. These will be described in detail in Chapter 5.

Support sustainable development. Sustainable development is generally considered equal in importance to social, economic, and environmental goals in formulating policy, which should ultimately meet the needs of the current generation without compromising the ability of future generations to meet their needs. To that end, much of adaptation is just good development policy. But how must development policy change because of climate change? Much evidence suggests that the societies most resilient to climate change are those with stable and effective institutions and high GDP. Therefore, sustainable development policy in the context of climate change requires measures that generate income through climate change mitigation activities or build the resiliency of agroecosystems to climatic perturbations. In particular, financial incentives and markets that encourage farmers to build resiliency through synergies between mitigation and adaptation that also raise incomes will be the backbone of longer-term measures to build resiliency. For the Asia and Pacific region, it will be important to give equal weight to climate change and underdevelopment, as much of the population remains poor and dependent on the agricultural sector as a source of livelihood.

Support measures that promote efficiency, productivity, and alternatives to agriculture. Investments in the agricultural sector are needed to raise productivity and efficiency, which will ultimately alleviate the dependence on agriculture as a source of livelihoods for many people vulnerable to climate change. These investments will include agronomic research, increased performance and productivity in irrigation and water use, and improvements in related infrastructure and equipment. For those farmers and their children that need to move out of food production, programs for retraining and employment opportunities need to be created. As a result, investments in education and retraining, green sector development, and social protection programs need to be bolstered.

1.6 Conclusion

Climate change will bring enormous challenges to the Asia and Pacific region. In this chapter, broad indicators have been presented in order to give a sense of the exposure, sensitivity and adaptive capacity of people in the region. A review of the indicators highlight the vulnerability of the agricultural sector as a livelihood source for many and source of food security for all of its inhabitants. The review also exposed the large heterogeneity in farming systems across Central Asia, East and Southeast Asia, South Asia and the Pacific Islands. Existing undernourishment, poverty, and slowing productivity growth put many at a disadvantage that will be exacerbated by

Page 18: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

18

climatic change. The review of the indicators also highlights the range of exposure and vulnerability to climate change across the region.

Global warming is expected to have multifaceted impacts on the Asia Pacific region. Overall, the region is expected to become warming, with a large degree of variability, depending on latitude. In general, northern regions will experience greater warming and than those of lower latitude. While the Pacific countries will experience the lowest mean annual changes in rainfall and temperature, sea level rise is expected to significantly alter livelihoods and livability on some of the smaller islands in particular. Coastal areas in South and Southeast Asia will face the triple threat of changing precipitation, temperature, and sea level rise. Finally, cooler northern regions are expected to warm, which may bring welcome news to farmers in terms of longer growing seasons.

The combination of poverty in rural areas combined with expected impacts from climate change and remaining uncertainty will require careful planning for adaptation. Scarce budgetary resources and important existing claims on these resources for crucial social development, such as education and health, as well as for emergency assistance further support careful targeting in the spirit of building resiliency against climate change, which calls for greater flexibility in decision-making and investments.

Page 19: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

19

CHAPTER 2. IMPACTS OF CLIMATE CHANGE ON FOOD SECURITY

Key Messages

Developing countries in Asia are likely to face the highest reductions in agricultural potential in the world due to climate change.

Scenarios consistently project that South Asia’s agriculture will be adversely affected by climate change, while East Asia’s agricultural capacity is likely to increase. Crops important for food security in South and Southeast Asia will be negatively affected.

o South Asian agriculture is likely to be one of the most adversely impacted in the world by climate change. In many cases, CO2 fertilization’s positive effects will not be enough to reverse declining yields.

o Global assessments project that East Asia is likely to benefit from climate change if the beneficial effects of CO2 fertilization are considered. Modeling studies show, however, that under scenarios of greater warming and lower precipitation, crop yields and livestock production in the region might be adversely impacted.

o Crops critical for food security are likely to be adversely affected in Southeast Asia, causing a decrease in human cereal consumption in the region.

o Any impact assessment needs to consider the high rates of land degradation in Central Asia, as well as possible increases in land conflicts.

o Climate change will affect specific crops and fisheries in the Pacific Island countries, with potential negative consequences for food security. Additional research will be needed, however, to obtain more specific results.

A warmer and drier climate will reduce the agricultural GDP of all countries in Asia, particularly in South and Southeast Asia. Economic losses in the Pacific Island countries are also likely to be high.

Projected increases in the frequency and intensity of climate extreme events are likely to substantially affect the agricultural GDP of the Asia and Pacific countries, leading to worsening trade conditions.

Food prices in developing countries will substantially rise as a result of climate change. Overall, climate will have a rather strong effect on cereal prices and a moderate impact on livestock prices; both cereal and meat prices will increase with climate change as compared to the no climate change scenario. Price increases for cereals and soybeans due to climate change are on the order of 20 to 70 percent by 2050 while maize crop, will be hit especially hard by climate change with prices more than doubling.

Women are more exposed to the impacts of climate change on agriculture.

Lack of secure land tenure might contribute to the displacement of poor farmers from marginal lands to be used in mitigation efforts, as well as cause land degradation, exacerbating global warming. Both land degradation and displacement can result in reduced food availability and fewer ways to access food.

Page 20: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

20

Climate change might increase the global burden of disease because more frequent and severe floods and droughts, as well as changes in mean temperature and rainfall, are likely to increase the number of people at risk.

Health impacts in developing countries will be mostly negative.

Vulnerable people who already face the highest burden of climate-sensitive diseases in the world will be the most affected.

Impacts of climate change will be highly dependent on the capacity of countries to limit transmission and treat infections.

The Asia and Pacific region is already highly prone to natural disasters. Climate change will intensify climate extreme events, increasing food insecurity with possible long-term consequences for lifetime earnings.

Climate change might intensify conflicts in Central and South Asia as a result of water scarcity and more intense climate extreme events.

2.1 Introduction

Climate change is already evident in a number of ways. Consistent warming trends and more frequent and intense extreme events have been observed across the Asia and Pacific region in recent decades. In line with that trend, climate change scenarios consistently project temperature increases across the Asia and Pacific region. There is much less certainty and agreement among models on rainfall variability and frequency and intensity of extreme climate events. It is generally expected, however, that extreme events such as cyclones, floods, hailstorms, and droughts will increase in frequency and severity across the region, as well as in specific areas. Such changes in climate are expected to have significant impacts on food security.

According to the FAO, a food system is vulnerable when one or more of the four components of food security—food availability, food accessibility, food utilization, and food system stability—is uncertain and insecure (FAO 2008b). Food availability refers to the physical quantities of food that are produced, stored, processed, distributed, and exchanged. Food accessibility refers to the ability to secure entitlements defined by a set of resources (such as legal, political, economic, social, and cultural) that an individual requires to obtain access to food. Food utilization is associated with the use of food and how a person is able to secure essential nutrients from the food consumed. Finally, food system stability relates to the temporal availability of, and access to, food (FAO 2008b).

Climate change will affect all dimensions of food security. Projected impacts on agricultural potential and crop yields will cause significant changes in cropping patterns. As a result, distribution and exchange of food among countries will be affected. Vulnerable populations, particularly women, are likely to be the most affected by increases in food prices and changes in income-generating capacity. Changes in patterns of climate-sensitive diseases and deterioration of food quality might affect people’s ability to make use of the nutritional value of foods. Furthermore, climate extreme events and conflicts might increase as climate change affects food system stability (FAO 2008b).

Page 21: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

21

Many other factors related to the four dimensions of food security are projected to be affected by climate change.4 This chapter does not comprehensively address every indicator affected in each dimension in the Asia and Pacific region, but it tackles important physical, economic, social, and political aspects that show that Asia and Pacific populations, particularly vulnerable ones, will be severely affected by climate change. For the first dimension—food availability—it addresses impacts on agricultural potential and crop yields. It also mentions impacts on agricultural GDP and trade to assess the distribution impacts of climate change. For the second dimension—food accessibility—the chapter assesses impacts on food prices, income, labor productivity, land tenure, and gender. Health impacts are evaluated as an indicator of impacts on food utilization. Finally, impacts on climate extremes and conflicts will be discussed in relation to food system stability. Climate extremes, however, will be mentioned throughout this chapter as expected increases in frequency and intensity of extreme events will affect all dimensions of food security.

The real impacts of climate change on agriculture and food security can be understood only if all dimensions of climate change are assessed in an integrated way. These dimensions of climate change include temperature and rainfall changes, sea-level rises, climate extremes, and water scarcity, and their several socioeconomic, political, and environmental implications. A good understanding of the impacts of climate change and variability on food security and agriculture is essential to improve adaptation and mitigation strategies and thus reduce vulnerabilities.

2.2 Food availability

Food availability is determined not only by a country’s capacity to produce and stock food, but also by its net food imports (FAO 2008b). The first part of this section addresses the impacts of changes in mean temperature and rainfall on agricultural potential and crop yields in Asia and Pacific regions and countries. The second part addresses impacts on food trade among countries. Because developing countries will grow more dependent on food imports, the impacts of climate change on countries’ agricultural GDP will also be assessed.

2.2.1 Impacts of climate change on agricultural potential and crop yields—literature review

Prospects for agricultural potential in Asian developing countries

Although parts of the developed world might benefit from global climate change, much of the developing world may experience a decrease in its food supplies and an increase in the risk of hunger and malnutrition, particularly in already food-insecure regions. According to several studies, production areas will shift from developing countries (particularly in the arid and subhumid tropics) to developed ones, increasing regional differences in crop production and cereal yields with significant polarization of effects (Rosenzweig et al. 2001; Parry et al. 2004; Parry, Rosenzweig, and Livermore 2005; Schmidhuber and Tubiello 2007; Easterling et al. 2007). Findings in the Fourth Assessment of the Intergovernmental Panel on Climate Change (IPCC) show that even small increases in temperature would decrease yields in lower-latitude countries, and more than two degrees of warming would decrease yields in almost all parts of the world (Parry 2007). Furthermore, favorable agricultural conditions due to yield and adaptation 4 For a full description of the food security framework and possible impacts of climate change on the four dimensions, go to

FAO (2008b).

Page 22: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

22

potential in developed countries are likely to aggravate inequalities in countries’ development potential (Parry 2007; Easterling et al. 2007).

Asian developing countries are likely to face the most significant negative impacts on cereal production. Scenarios show that by 2080, cereal production in Asia is expected to decline approximately 4 to 10 percent compared with scenarios without climate change for the same year (Figure 2.1) (Fischer, Shah, and Van Velthuizen 2002; Fischer et al. 2005). As a result, Asian developing countries will be responsible for two-thirds of a 2–4 percent reduction in human cereal consumption in developing countries projected to happen in 2080 (Fischer et al. 2005).

Such decreases in food supply can have serious consequences for national food security. Because of projected climate change and increases in the number of poor people, scenarios consistently predict that climate change will increase the number of people at risk of hunger in comparison with reference scenarios (without climate change) in countries in South Asia, which is second only to Sub-Saharan Africa (Parry, Rosenzweig, and Livermore 2005; Parry et al. 2004; Easterling et al. 2007; Lobell et al. 2008).

Climate change will adversely impact agriculture in all regions in Asia if the beneficial effects of CO2 on plants are not considered (Parry et al. 2004; Cline 2007). If those effects are fully realized, East Asia’s agriculture is likely to benefit from the combined effect of global warming and CO2 fertilization, whereas crops in South Asia are still likely to be harmed. In the other regions—Southeast Asia, Central Asia, and also in the Pacific Island countries—there is either substantial disagreement among scenarios on likely outcomes or there is not enough research on the matter (see Cline 2007; Parry et al. 2004; Fischer 2002). Information about model categories can be found in Annex 2.1. Most important, it is still unknown if the expected beneficial effects of CO2 fertilization will actually happen, as many experiments are conducted in closed chambers that do not realistically replicate all of the stresses that plants suffer on farms (see Box 2.1). Therefore, experiments might be underestimating the effects of global warming on agricultural potential. More research will be needed on CO2 fertilization and its likely effects on crops.

Projections from modeling studies show that crops important for food-insecure populations in South and Southeast Asia will be negatively impacted by climate change (Fischer et al. 2005; Lobell et al. 2008). Simulations project that the regions of South and Southeast Asia will face the largest decreases in wheat production in the world (20–75 percent and 10–95 percent declines, respectively), and Southeast Asia will have substantial decreases in attainable rice production (Fischer, Shah, and Van Velthuizen 2002; Fischer et al. 2005; Lobell et al. 2008).

Many studies have assessed the impacts of climate change on agriculture around the world. The remainder of this section summarizes some of these studies for the Asia and Pacific countries (Table 2.1 and 2.2 present a summary of the results of crop yield and livestock studies). Each section begins with a summary of global assessments results for each region, followed by crop and livestock studies. Disagreements between global assessments and case studies over the direction and magnitude of impacts illustrate the importance of investigating impacts at the local level and considering context-specific factors.

South Asia

All countries in South Asia are expected to see decreases in agricultural production capacity, even considering the effects of CO2 fertilization on crop yields (Figure 2.2). Without sufficient

Page 23: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

23

adaptation measures, several South Asian crops important to large food-insecure populations will be affected by climate change. In tropical parts of South Asia, rice and wheat are already being cultivated close to their temperature tolerance. A study that uses 20 general circulation models to analyze climate risk in 12 food-insecure regions shows that 95 percent of climate models agree that by 2030, wheat crops in South Asia will be harmed by climate change. At least half of the models also projected production loss in rapeseed crops greater than 5 percent (Lobell et al. 2008).

India. By 2080, India is expected to lose around 18 percent of its rainfed cereal production potential (Fischer et al. 2001). Other projections estimate losses in average output potential (considering irrigation and rainfed agriculture) of around 30 percent with CO2 fertilization and 37 percent without (average of Northeast, Northwest, Southeast, and Southwest regions) (Cline 2007). By 2080, scenarios also consistently project losses in crop yield potential (Parry et al. 2004). Another study shows that even small temperature increases will reduce rice yields if the beneficial effects of CO2 fertilization are not considered. If temperature increases are less than 4oC, the inclusion of CO2 effects in scenarios results in increases in rice yields (Aggarwal and Mall 2002; Saseendran et al. 2000).

Bangladesh. In Bangladesh, scenarios project that rice yields are likely to increase under a warming climate with atmospheric CO2 concentration at 580 or 660 parts per million (ppmv) whereas wheat production is likely to decrease under most scenarios. A study investigates the impacts of climate change on three varieties of rice (aus, aman, and boro) and wheat yields using CERES-Rice and CERES-wheat models (Karim, Hussain, and Ahmed 1999). With a CO2 concentration of 330 ppmv and temperature increases of 2oC and 4oC, potential production decreases for all crops and varieties. With a CO2 concentration of 580 ppmv and the same temperature increases, rice production increases (except for aus under a 4oC scenario). Under the same conditions, wheat production decreases by 1 and 40 percent (2oC and 4oC, respectively). Finally, for 660 ppmv of CO2, production increases for all crops, except under a 4oC scenario for wheat, which faces a decrease of 31 percent. Therefore, while rice crops are likely to benefit from the combined effect of CO2 fertilization and global warming, wheat production is likely to be severely affected.

Another study that combines the impacts of moisture stress, inundation, and salinity from global warming projects that with a CO2 concentration of 660 ppm and temperature increases of 4oC, rice and wheat production will decrease by an average of 8 and 32 percent, respectively (Faisal and Parveen 2004).

Pakistan. A study in Pakistan shows that a small temperature increase of 0.9oC will likely increase wheat yields in humid and semi-arid regions (assuming beneficial effects of CO2 on crops). A 1.8oC temperature increase, however, is likely to decrease wheat yields in a semi-arid zone (Faisalabad), even taking CO2 effects into consideration (O'Brien 2000). Like the results for other countries, results for Pakistan show that impacts will vary depending on the natural zone and projected temperature.

Afghanistan. Results of seven HadCM3 scenarios suggest that Afghanistan is the country expected to face the most consistent crop yield reductions in Asia, ranging from 5 to 30 percent (Parry et al. 2004).

Page 24: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

24

East Asia

For East Asia, at the regional level, there seems to be some consensus that CO2 effects will outweigh the adverse effects of global warming by the 2080s (Parry et al. 2004; Cline 2007; Fischer, Shah, and Van Velthuizen 2002). The studies summarized here, however, find that climate change’s impacts on crop yields and livestock production are highly dependent on region, CO2 fertilization level, intensity of warming, and direction of precipitation and, therefore, highly dependent on the assumptions underlying the scenarios.

China. Most climate scenarios find that China’s agriculture is expected to benefit from climate change (Fischer, Shah, and Van Velthuizen 2002; Fischer et al. 2005; Parry et al. 2004; Cline 2007; Fischer et al. 2001). Four out of five Agroecological Zone (AEZ) models predict increases in cereal-production potential in the range of 5 to 23 percent (Fischer et al. 2005). Another study shows that by 2080, agricultural production capacity will increase in all regions, except for the Northwest and Yellow Sea, if CO2 effects are considered (Figure 2.3). If the benefits of carbon fertilization are not considered, however, all regions (except for the South Central region) face decreases in production capacity (Cline 2007).

The positive effects of CO2 fertilization will considerably reduce the negative impacts of climate change on crop yields in China, according to a study that assesses the impacts of temperature increases on irrigated and rainfed rice, maize, and wheat yields (Erda et al. 2005). By the 2080s, for a medium-low emissions scenario B2 (temperature and rainfall increases of 3.20oC and 10.2 percent), irrigated and rainfed wheat yields are likely to face substantial increases. Rainfed maize yields are also expected to increase, while irrigated maize yields will decrease. Both irrigated and rainfed rice yields, however, are expected to decrease. For the other scenario investigated, a medium-high emissions scenario A2 (temperature and rainfall increases of 3.89oC and 12.9 percent), all crops are likely to face yield increases with CO2 fertilization, except for irrigated maize. Without CO2 fertilization, however, all crop yields are projected to decrease. The study, however, did not consider future availability of water for irrigation and more pessimistic scenarios regarding changes in rainfall. These impacts should be further considered.

Mongolia. In Mongolia, livestock production, which is the only source of income for more than 34 percent of the population, is expected to be adversely impacted by global warming through reductions in pasture biomass and animal weight (Batima 2006; Shagdar 2002). More than 80 percent of the total land area in the country is used for agriculture, of which 97.5 percent is used for pasture. Also, more than 90 percent of Mongolian livestock’s annual feed intake comes from pasture (Batima 2006). Furthermore, the Mongolian economy relies heavily on livestock production; its derivatives account for one-half of the country’s output and almost 90 percent of its exports (Bolortsetseg and Tuvaansuren 1996). Therefore, climate change impacts on pasture directly affect livestock weight and thus the livestock industry. Impacts on pasture will vary according to the region. Global climate modeling (GCM) scenarios suggest that a warmer and drier climate will reduce pasture productivity in steppe and forest steppe and will reduce pasture quality in all natural zones of the country (Batima 2006). By 2080, estimated ewe weight is expected to decrease in the summer-autumn season by 0.18 to 7.75 percent, depending on the scenario and natural zone (Batima 2006). Steer weight is also expected to be affected by climate change, with estimated reductions of approximately 15–30 kilograms (kg) by the end of the autumn (Bolortsetseg and Tuvaansuren 1996).

Page 25: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

25

Batima’s study shows that over the past 40 years, the peak of pasture biomass has declined by 20–30 percent. During recent decades, the weights of ewes, goats, and cattle have decreased by 3.63 kg, 2.0 kg, and 13.8 kg, respectively. One of the reasons is the difficulty of grazing during the summer because of higher air temperatures. The threshold temperatures for grazing are 16–19oC in high mountains, 20–22oC in the steppe, and 26oC in the Gobi Desert. Therefore, even in regions where aboveground biomass is expected to increase (high mountains and desert steppe), increased temperatures might adversely affect livestock production.

Republic of Korea. In the Republic of Korea, global warming is likely to decrease yields of rice, the country’s major crop. According to one study (using ORYZA 2000), however, the beneficial effects of CO2 fertilization are likely to outweigh a temperature increase of 1oC, reversing an expected decrease in crop yields of 7.5 percent (without CO2 effects) to an increase of 4.9 percent (with doubled CO2 concentration). For higher temperatures, the effects of CO2 fertilization are not enough to reverse the declining yields (Kang Su and Chung Kuen 2006).

A document from the Government of the Republic of Korea finds that a 3–4oC temperature increase is likely to expand suitable land northward and increase crop diversification. If temperature rises excessively, however, perennial temperate fruit trees such as apple trees might be harmed in some regions in the south. In the southern coastal regions, pears, peaches, grapes, sweet persimmons, tangerines, citrus, and kiwi are likely to benefit from warming. Finally, global warming is likely to push cultivation of cool-season vegetables from the highlands to even higher lands or further north (Government of the Republic of Korea 2003).

Southeast Asia

In Southeast Asia, a recent study projects losses in agricultural production capacity for all countries by 2080, even if CO2 fertilization is considered (no data are available for Lao PDR and Singapore) (Cline 2007). Reductions vary from an estimated 39.3 percent in Myanmar to 15.1 percent in Viet Nam without the effects of CO2 fertilization, and from 30.1 percent to 2 percent, respectively, with CO2 effects (Figure 2.4) (Cline 2007). Another study, however, predict small positive and negative variations in crop yields depending on scenario and country (Parry et al. 2004).

There seems to be consensus that important crops for food security in the region will be negatively affected by climate change, despite disagreements among modeling studies about impacts on agriculture as a whole (Fischer, Shah, and Van Velthuizen 2002; Lobell et al. 2008). One study shows that 95 percent of models (total of 20 GCM models) project losses in rice yields in Southeast Asia as a result of climate change (Lobell et al. 2008). This finding is a reason for concern, as the region is one of the most dependent on rice for daily calories in the world (Nguyen 2005). Another study predicts that Southeast Asia is likely to see its human cereal consumption decrease by 0.4 to 4.2 percent, depending on the scenario (Figure 2.5). As the figure shows, even in optimistic scenarios in which the world human cereal consumption is expected to increase, cereal consumption in Southeast Asia is still expected to decrease in comparison with scenarios without climate change.

The Philippines. A study in the Philippines verifies the negative relationship between global warming and crop yield, finding that between 1992 and 2003, for each 1oC increase in the minimum temperature in the dry season, rice yield declined by 10 percent (Peng et al. 2004). The study used simulation models to evaluate the impact of temperature increases on rice yields and

Page 26: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

26

analyzed weather data (1979–2003) and data from irrigated field experiments (1992–2003) conducted at the International Rice Research Institute (IRRI) farm in the Philippines.

Another study by the Ministry of Environment of the Philippines that assumes a double CO2 concentration (compared with 1995) projects changes in rice yields from -85 to 23 percent depending on scenario, variety, cropping season, and study site (CCCM, GFDL, UKMO, GISS scenarios). For corn yields, results were overwhelmingly negative in all situations and scenarios (DENR 1999).

Indonesia. The production of paddy in a crop year will be reduced by 1.3 million tons with a 1oC change in the August Niño 3.4 SSTA (sea-surface temperature anomalies of the Central Pacific Ocean), based on econometric models (Falcon et al. 2004).

According to Naylor et al. (2007), estimating changes in precipitation for Indonesia is particularly difficult given that the archipelago is situated in an especially sensitive region where precipitation is influenced by changes in both mean temperature and the large-scale dynamical circulation in the tropics. Moreover, the coarse grid sizes (typically 200 _ 200 km2) of GCMs do not resolve the regional-scale (e.g., 50 _ 50 km2) interactions between the large-scale atmospheric circulation and the very complex and mountainous topography of the archipelago; these interactions are important contributors to the hydrological cycle over Indonesia. The islands of Java and Bali are not even represented as land in many GCMs.

Using empirical downscaling models for Java and Bali, the authors find that total rainfall is expected to increase in AMJ relative to the current pattern, but decrease in JAS. In AMJ, total rainfall is projected to increase by 10 percent in the study regions. In JAS, however, nearly all models project a decline in rainfall. Total rainfall is projected to decline by 10–25 percent on average and by as much as 50 percent in West/Central Java and 75 percent in East Java/Bali at the tail end of the distributions. Some models project that total rainfall will drop close to zero for the JAS season. The predicted increase in AMJ rainfall would not compensate for reduced rainfall later in the crop year, particularly if water storage for agriculture was inadequate. Second, the extraordinarily dry conditions in JAS could preclude the planting of rice and all other crops without irrigation during these months by 2050. Temperature is expected to increase under global warming, and impacts on rice are expected to be negative for Indonesia, following the results from IRRI, suggesting that rice yields are closely linked to mean minimum temperatures during the dry season; for every 1°C increase in the minimum temperature, rice yields decrease by 10percent.

Mekong Region. Around 10,000 square kilometers (km2) of the Mekong River Delta, or one-fourth of its total area, are under rice cultivation, making the area one of the major rice-growing regions of the world (Nguyen 2007). With a one-meter sea-level rise, half of its mangrove area (2,500 km2) will be lost and around 1,000 km2 of cultivated farmland and aquaculture area will become salt marshes (Cruz et al. 2007).

In the Mekong River basin, rainfed rice cultivation, the main source of food in the region, will be adversely affected by hydrological change caused by global warming (Chinvanno 2003). One study assesses the impacts of climate change on rainfed rice cultivation in study sites in Lao PDR, Thailand, and Viet Nam (lower Mekong River basin). Results show that rice yields in the sites located in Lao PDR and Viet Nam (summer-autumn crop) are expected to be negatively affected by climate change. On the other hand, projections suggest that the Ubonratchathani

Page 27: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

27

Province in Thailand is likely to benefit from a warmer climate. The study did not, however, assess the impacts of more frequent and intense floods (Snidvongs 2006), so there is room here for further research.

Central Asia

According to some studies, countries in Central Asia are likely to increase their agricultural production capacity (Cline 2007; Fischer et al. 2001). Kazakhstan in particular is likely to benefit in scenarios with and without CO2 fertilization effects, whereas other countries might benefit only if CO2 fertilization effects are considered (Figure 2.6). One study projects that most gains in rainfed cereal production potential in the region will be in Azerbaijan, Kazakhstan, and Turkmenistan (85 percent of the increase). Georgia is the only country likely to face decreases in production potential (Fischer et al. 2001). According to another study, however, Central Asia is expected to lose between 5 and 10 percent in crop yield potential, even considering CO2 fertilization effects (Parry et al. 2004). Therefore, it is reasonable to say that it is uncertain whether growing conditions will deteriorate or improve in Central Asia as a result of climate change (Pandya-Lorch and Rosegrant 2000), and more research will be needed on this topic.

The impacts of climate change on agriculture in Central Asia are interwoven with its impacts on water resources. The region is arid, and more than 70 percent of its agricultural land requires irrigation (Schubert et al. 2008). As discussed later in this chapter, climate change is likely to aggravate conflicts over water resources, given the water interdependence of countries in the region (Perelet 2007). Therefore, although agricultural capacity is expected to increase, conflicts over scarce water resources are likely to reduce those benefits. The Fergana basin, situated in Kyrgyzstan, Tajikistan, and Uzbekistan is particularly vulnerable to climate change and prone to conflicts, because it is the most important area of agricultural cultivation and the most densely populated part of the region (Schubert et al. 2008). Furthermore, climate change may aggravate soil degradation and salinization of large areas of Central Asia as a result of poorly managed agricultural irrigation, clearing of forests, overgrazing, and unsustainable agricultural practices (Schubert et al. 2008).

In the national communication reports for the United Nations Framework Convention on Climate Change (UNFCCC), governments of Central Asian countries reported how they expect the agriculture and livestock sectors in their countries to be affected by climate change. The reports reveal that climate change might benefit several crops such as heat-tolerant grapes in Armenia; potato, maize, tobacco, and sugar beet in Georgia; cotton, tobacco, and sugar beet in Kyrgyzstan; and cotton and irrigated cereals in Uzbekistan. Adverse impacts are expected to be found in spring wheat and wool productivity in Kazakhstan; livestock in Kyrgyzstan; and potato, horticulture, and livestock in Armenia.5

Pacific Island countries

Climate change is already showing its effects in the Pacific Islands. In some islands and atolls in the Federated States of Micronesia (FSM), Marshall Islands (RMI), and Palau, taro pits have been contaminated by saltwater owing to depleted freshwater lenses, extended droughts, and saltwater intrusion (Shea 2001). Taro and arrowroot crops in the Marshall Islands already show

5 The national communications of all non–Annex I countries are available at the UNFCCC website:

http://unfccc.int/national_reports/non-annex_i_natcom/items/2979.php.

Page 28: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

28

signs of stress under present conditions and might not survive further increases in temperature (FAO 2008a). In the Cook Islands, global warming has been associated with increases in insect populations, which are killing domestic pigs on one island in the Northern Group. As a result of drier lands, taro growth has been constrained in Mangaia, Aitutaki, Pukapuka, and Mauke (FAO 2008a).

Table 2.2 shows projected impacts of climate change and variability on crop yields in Viti Levu (Fiji’s largest island). Cassava is expected to be adversely affected by future changes in climate in all scenarios, whereas impacts on taro and yam crops depend on the scenario. Projections suggest that in Viti Levu, an 8 percent increase in rainfall, expected by 2050, would benefit most crops except yam. A decrease by the same amount would hurt most crops, particularly sugarcane, which accounts for 45 percent of Fiji’s exports and is cultivated primarily in Viti Levu (World Bank 2000). For the low island Tarawa in Kiribati, climate change is expected to affect agricultural crops through changes in rainfall, increases in droughts, and sea-level rise. Increases in rainfall will benefit coconut, breadfruit, and te babai crops, while decreases will harm coconut and te babai crops (World Bank 2000).

In the Pacific Island countries, food security is highly correlated with the fisheries industry. Many Pacific Islands depend on coastal marine sources for protein for daily consumption. Most of the production (around 80 percent) does not enter the cash economy. Some of the most developed and agriculturally oriented countries consume more than 50 kg of seafood per capita annually, while some remote atolls consume more than 250 kg (FAO 1997). Climate change will affect oceanic and coastal fisheries by changing coastal circulation patterns because of sea-level rise and global warming. As a result, these changes may affect nutrient supply, lagoon flushing, coastal erosion, ocean acidity, and coral bleaching, which in turn will affect the reef-building capacity of corals and the spawning cycles of reef fishes and invertebrates (FAO 2008a).

Box 2.1. The importance of understanding assumptions behind the scenarios

Disagreements among modeling studies with regard to the future impacts of climate change on agricultural capacity and crop yields are in part a result of different assumptions, some of which are described here.

Using crop models to derive yield functions has high levels of uncertainty, especially when considering the effects of CO2 fertilization. The most common crop models are also those that have been evaluated the least using available data from elevated CO2 experiments (Zhu 2007). Experimental data used in climate simulations might lead to larger estimated crop responses to elevated CO2 than will actually occur at the farm level (Fischer et al. 2005; Parry, Rosenzweig, and Livermore 2005). Weeds, diseases, and insect pests are assumed to be controlled, and no problems with soil conditions (such as salinity or acidity) or acid deposition are considered (Parry et al. 2004). Hence, model projections of future yields should use much lower CO2 fertilization factors (Zhu 2007).

Temperatures near the upper limit for crops can depress yields irrespective of CO2 concentration. Studies have demonstrated that yield responses to CO2 depend on temperatures (Zhu 2007).

Climate extreme events and other stressors—such as increased climate variability, sea-level rise, and land degradation—are often partially or entirely ignored (Fischer et al. 2005; Parry et al. 2004). Yet the impacts of increased incidence of natural disasters in agricultural systems might be even more severe than the impacts of increases in mean temperature and precipitation (Easterling et al. 2007). Water-modeling processes also suffer from significant uncertainties, mainly because of the uncertainty in

Page 29: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

29

precipitation inputs (Kundzewicz et al. 2007; Arnell 2003), which need to be better understood.

Global models applied at the regional level may not represent the variability of agricultural systems within similar agroecological zones or dissimilar agricultural regions (Parry et al. 2004). Small-scale elements might not be properly represented, such as warm and cold fronts, hurricanes, and diversity of ecosystems and land use (Parry et al. 2004). Modeling of hydrological processes, including infiltration, runoff, and evaporation, are highly simplified. Precipitation is poorly represented in GCM results (Parry et al. 2004).

Analyses of farm-level adaptation strategies simulate only the range of agricultural technologies currently available. Because they do not incorporate future agricultural technology, they potentially overestimate impacts of climate change (Parry et al. 2004).

Changes in production functions are often not considered. Studies show that the amount and distribution of arable land and crop suitability will be altered. Climate change might result in more land being allocated to livestock production, and responses to nitrogen fertilization may be altered because of changing nutrient solubility in warmer soils (Zhu 2007; Parry et al. 2004).

The future impacts of climate change on agriculture are highly dependent on projected socioeconomic scenarios. More globalized world scenarios (A1F1 and B1) experience greater reductions in crop yield than the scenarios of a more regionalized world (A2 and B2) (Zhu 2007). There is also uncertainty about the magnitude of climate change and its spatial and temporal distribution, particularly regarding future precipitation changes (Lobell et al. 2008; Fischer et al. 2005).

Quality and reliability of data sets are uneven across regions (Fischer et al. 2001).

2.2.2 Impacts of climate change on agriculture and crop yields—combined IFPRI crop modeling, neural network and IMPACT results

The IFPRI modeling effort has three main parts – detailed crop modeling for 5 key crops (rice, wheat, maize, soybeans, and groundnuts), estimation of a nonlinear reduced form function for each crop variety that incorporates a wide range of biophysical and climate drivers, and the IMPACT model for projections of world agricultural production, consumption and trade. Details on each element are contained in Annex 2.2 on the modeling environment.

An important consideration is which of the many available future climate scenarios to use to estimate the effects of climate change on agriculture. We base our results on the Worldclim downscaling of the Hadley GCM output for the SRES A2a scenario, available at http://www.worldclim.org/. It is important to stress that this is a single realization of climate from a single model. The implication of this assumption can be seen in Figures 2.7 and 2.8 comparing the Hadley with the GCM3.1 (T63) model. Note the substantial differences in both temperature and precipitation in Asia and elsewhere. Both GCMs use identical GHG emissions but have significantly different regional climate outcomes (see Figures 2.7 and 2.8). Simulations with additional GCMs will be done.

Crop modeling results

Crops, and indeed individual varieties of crops, have different responses to the three main elements of climate change – precipitation, temperature and atmospheric CO2 levels. Our final results incorporate detailed variety responses using climate and soil information from spatial data sets with 5 arc minute resolution (approximately 10 km at the equator). However, it is useful to look at some summary response measures. Figure 2.9Figure 2.9 and Figure 2.10 show rice and

Page 30: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

30

maize average responses to different levels of temperature and nitrogen applied, under both rainfed and irrigated conditions. The temperature range of – 5 degrees to + 5 degrees around 2000 temperatures is much larger than is expected by 2050. As expected, higher temperatures cause lower yields generally. But an important result to note is that higher temperatures have much less effect on yields with low levels of N application.

The all-Asia average results conceal significant regional differences. Figure 2.11 and Figure 2.12 report the same response combinations for rice but confined to China in the case of Figure 2.11 and India in the case of Figure 2.12. Note the very different country averages. For China, the average effect of an increase in temperature is uniformly positive across different combinations of temperature and nitrogen application. In India, on the other hand, higher temperatures are almost uniformly negative.

Given that there is some agreement among the crop modeling community that DSSAT overstates the CO2 productivity effect--version 4.5 reportedly has smaller effects but it has not yet been released—we used a CO2 level of 369 ppm. Figure 2.13 and Figure 2.14 report all Asia average CO2 fertilization effects for rice and maize respectively. As is commonly reported, higher levels of atmospheric CO2 concentrations do result in higher yields, but importantly this effect is limited to higher levels of N application. With low levels of N, the CO2 fertilization effect doesn’t exist. And in fact, in some circumstances, such as maize with moderate N application levels, it can be negative. Essentially what happens is that any productivity effect is with respect to growth of plant biomass other than seeds.

The Crop Model-Neural Net Output

We take outputs from several million individual runs of the crop model to estimate a feedforward neural net model (a nonlinear regression technique). This allows us to use location-specific information (at 30 arc second resolution) to simulate the effects of climate change on crops throughout Asia, including changes in the cropping calendar. The major steps included in the neural network analysis are crop simulation modeling, translation to a neural network approximation, and aggregation of the neural network predictions to regional levels suitable for incorporation into IMPACT. To allow multiple simulations of climate effects for the entire surface of the globe, we developed a reduced form implementation. We ran the crop model for each crop and variety with a wide range of climate and agronomic inputs and then estimated a feed-forward neural net (NN) for each of the 27 soil categories. We obtain a continuous and differentiable approximation of the crop model results that allows us to find the maximum possible yield and corresponding nitrogen input needed based on location-specific geophysical characteristics and climate. The results of this estimation process are fed into the IMPACT model. Further assessment of fit of neural net estimation algorithms and upscaling techniques from crop models to neural net, and to IMPACT will be done in finalizing results for the final report.

Figure 2.15 show a series of maps for yield changes of rainfed and irrigated rice, rainfed wheat and rainfed maize to climate change between 2000 and 2050 for various Asia regions. The climate changes are taken from the Hadley GCM using the A2a SRES scenario with the results downscaled to 30 arc second resolution, available at the www.worldclim.org. We also use these climate data to adjust the cropping calendar. In the maps, areas in green are where a crop can be grown in 2050 but not in 2000; areas in brown are where it can be grown in 2000 but not in 2050.

Page 31: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

31

Area, Yield, and Production Changes from IMPACT

The biophysical impacts of climate change from the crop model-neural net analysis are then input into the IMPACT model based on a classification of different regions as high-input versus low-input rainfed and irrigated areas, respectively, and represented as shifts in rates of crop yield and yield growth over time. In a further update for the final report, we plan to incorporate neutral network generated weighting factors for high and low input production for rainfed and irrigated crops. IMPACT then models area, yield, production, trade, and price data annually for 2000-2050 at levels that clear world market prices.

Crop calendar changes due to climate change obtained from the neural network process can lead to area go completely out of cultivation as well as area formerly not cultivated for a specific crop becoming available for cultivation. When the crop calendar in an area changes so that a crop that was grown in 2000 can no longer be grown in 2050, harvested area is gradually set to zero. However, when it becomes possible to grow a crop in 2050 where it could not be grown in 2000, we do not add this new area at this point. Additional details are provided in Annex 2.2.

Impacts from temperature and rainfall change on the production of key staple crops for the Asia and Pacific region are presented in Tables 2.3 to 2.5. Results exclude the Pacific Islands. Area, yield and production impacts are the result of both the biophysical shock from changes in temperature and precipitation, and the impact of commodity price changes induced by the biophysical shocks. As described in more detail below, the climate change shocks on production induce significant increases in prices for the key cereals, which in turn induce a supply response that dampens the initial climate change shock. With the net effect of these forces, maize production is expected to decline in Central, South and East Asia, whereas production is expected to be higher in the Southeast Asia region. Production changes are driven by changes in area and yield. Central and South Asia, particularly India, are expected to experience large declines in area harvested for maize under climate change as the growing period window for the crop is altered significantly. At the same time, however, Central Asia is projected to experience an increase in maize yield by 2050 under climate change while yields are expected to decline in East Asia and South Asia, particularly India.

Rice yield is expected to increase—or at least stay about the same—across all Asian developing countries examined in the climate change case compared to the baseline. The expected rice yield increase is particularly large in India but this is offset by a fairly strong decline in harvested area. This is due to an improvement in the growing climate for much of the upland rice area in India, small positive and negative yield changes across many irrigated production areas, and important yield improvements in the Chotanagpui, Godavari, and Krishna river basins. At the same time, however, significant area in India will become unsuitable for rice production, essentially cancelling out the increase in rice yield in other parts of the country, particularly the Brahmaputra and Brahmari river basins. Area suitable for rice production is also expected to decline in Central Asia and Southeast Asia, whereas higher temperatures and increased precipitation open up some new growing windows in for irrigated production in central-east of China, pushing up rice production by 10 percent compared to the no climate change simulation. Asia dominates the globe for total rice production (Figure 2.16), but total production will decrease in East Asia due to climate change while it will increase in South and Southeast Asia by 2050.

Page 32: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

32

Wheat yields are projected to decline in Central Asia (Figure 2.15) whereas production increases in both East and South Asia.

2.2.3 Impacts of climate change on trade and GDP—literature review

Impacts of a warmer and drier climate

Scenarios suggest that most of the economic damage in the global agricultural sector caused by climate change will occur in Asian developing countries (Mendelsohn and Williams 2004). The expected impacts on countries’ GDP depend on scenario and region. Poor countries, however, are likely to suffer damage in all scenarios. One study projects that by 2100, the poorest quartile of the world’s countries (54 countries, including Bangladesh, Bhutan, Lao PDR, Nepal, Pakistan, Viet Nam, and parts of India) will experience reduced GDP under all scenarios, whereas the richest quartile will benefit in all but one scenario (Table 2.6) (Mendelsohn 2006).

Many of those developing countries do not have the adaptation potential and coping capacities needed to deal with such losses, as shown by their current vulnerability to extreme events. Therefore, countries facing decreases in agricultural potential are projected to increase their net food imports by an additional 10 to 40 percent as a result of climate change by 2080, based on AEZ simulations (Fischer et al. 2005). The impact of climate change on food availability will thus also depend on the reliability of import capacity.

Regional assessments mask significant differences between countries. According to a modeling study, Asia, the former Soviet Union, and North America will benefit the most from climate change as a percentage of GDP (Mendelsohn et al. 2000). In contrast, other studies show that whereas the former Soviet Union and Eastern Europe will have large agricultural gains, the greatest damage will occur in the tropical regions of Asia, where agricultural GDP is expected to fall by 4 percent (Mendelsohn and Williams 2004; Fischer et al. 2005). Although China is expected to enjoy agricultural benefits of US$39 to US$65 billion from a 2oC increase, India is expected to suffer agricultural losses of up to US$86 billion and a decrease in net agricultural revenue of 55–91 percent, depending on the region and scenario (Mendelsohn et al. 2000; Cline 2007).

Alarmingly, the combination of a warmer and drier climate will have catastrophic consequences for the economies of Asian countries, although results are highly dependent on assumptions about temperature and precipitation changes. Another modeling study uses a global impact model (GIM) to evaluate the impacts of changes in climate on agricultural GDP according to different climate models (CCC, CCSR, PCM) (Mendelsohn 2005). Changes are then evaluated with two climate response functions (cross-sectional response and experimental climate response function). Results show that Bangladesh, India, and Thailand are likely to suffer damage in most scenarios. In general, however, results are highly dependent on scenario. A mild and wet warming (PCM) will increase agricultural revenues in several countries, while a hotter and drier climate (CCC) will harm all countries. In a hotter climate with just a slight increase in precipitation levels, most countries will still be harmed (CCSR).

For Southeast Asia, scenarios project a consistent reduction in agricultural GDP as a result of a changing climate. As shown in Figure 2.28, more pessimistic scenarios project losses of up to 5 percent of agricultural GDP in comparison with scenarios without climate change (Fischer, Shah, and Van Velthuizen 2002).

Page 33: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

33

In the Pacific Islands, without considering adaptation measures, total economic losses in Viti Levu (Fiji’s largest island) could be between US$23 and US$52 million by 2050, with sugarcane suffering the most significant damage. In the Tarawa atoll in Kiribati, average annual economic damage was estimated to be US$8–$16 million by 2050 (World Bank 2000).

As will be shown later in this chapter, losses can be substantially higher in years with severe climate events. Higher incidences of extreme climate events are likely to increase damage during El Niño years. A drought comparable to the one in 1997/98 could cost US$70 million in lost crops in Viti Levu. On the same island, an average cyclone could cause damage of more than US$40 million (World Bank 2000).

Impacts of more frequent, intense climate extreme events

Although the economic impacts of current and past climate change on agriculture in the Asia and Pacific countries are not fully understood, these countries have long known the economic impacts of natural disasters. For instance, a disaster risk assessment shows that climate extremes—especially floods—have a significant impact on the economies of Asian countries and that areas in South and Southeast Asia and parts of China are among the regions with the world’s highest risk of suffering economic losses from floods (Figure 2.29) (Dilley et al. 2005). Therefore, the projected increases in the frequency and intensity of climate extremes are likely to substantially damage the economies of the Asia and Pacific countries, taking into account current adverse effects of climate variability, which are briefly reviewed in this section.

In 2007, among the 10 countries with the highest economic losses (absolute losses in purchasing power parity [PPP]) as a result of extreme events, five were Asian countries. China ranked first (US$17,333 million), followed by the United States (US$12,366 million) and Bangladesh (US$ 10,774 million). Indonesia, Pakistan, and India were also in the list. For the period 1998–2007, China was first in terms of average losses (PPP), followed by the United States, India, and Bangladesh. Regarding average losses as a percentage of GDP in 2007, Tajikistan ranked first (10.44 percent) and Bangladesh ranked sixth (5.48 percent). Bangladesh was one of the countries most affected by climate extreme events in 2007. The vulnerability of Bangladesh to climate extreme events is also shown by its extremely high death rate, which was the highest for 2007 (2.98 killed per 100,000 people) (Harmeling 2008).

Examples of the adverse impacts of climate variability and natural disasters on the economy can be found in many countries of the Asia and Pacific region. For instance, in 1987 weak monsoon rains caused large shortfalls in crop production in Bangladesh, India, and Pakistan, contributing to a return to wheat importation in India and Pakistan (Parry et al. 1999). In another example, between 1978 and 2003, the average annual drought-affected area in China was estimated to be 14 million hectares, with an estimated direct economic cost of 0.5–3.3 percent of agricultural sector GDP (Pandey et al. 2007).

During drought years in the period 1970–2002, the ratio of loss to average value of total production was 3 percent in southern China, 10 percent in northeast Thailand, and 36 percent in eastern India (in China and Thailand, values were estimated only for rice; in India, values accounted for rice and nonrice crops) (Pandey et al. 2007). In absolute terms, production loss in India was estimated at US$856 million.

Furthermore, natural disasters affect the economy by affecting default rates. For instance, the 2005 Pakistan earthquake influenced the behavior of micro-borrower households. Delay in

Page 34: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

34

repayment was 52 percent higher in affected areas than in unaffected areas (Khan and Kurosaki 2007). Extreme events might also increase countries’ demand for imports, as well as foreign aid, with direct consequences for food availability.

Box 2.2 Vulnerability of Pacific Island countries’ economies to climate extreme events

Pacific Island countries have faced substantial economic losses as a result of natural disasters in recent decades. In the 1990s, the cost of extreme events in the region was estimated to exceed US$1 billion. During the 1997/98 El Niño event, Fiji’s losses in the sugarcane industry were around FJ$104 million, and agriculture losses associated with livestock deaths were around FJ$15 million (FAO 2008a). Total damage amounted to US$140–165 million, equivalent to 10 percent of Fiji’s GDP (World Bank 2000). As a result of Cyclone Gene in February 2008, Fiji experienced damage worth FJ$45 million in agriculture (excluding the sugar industry), infrastructure, utilities, and property. After that cyclone, Fiji’s government had to provide FJ$1.7 million worth of food rations (FAO 2008a). In 2003, Cyclone Ami caused US$35 million in lost crops in Fiji; in 2004 heavy flooding of the Wainibuka and Rewa Rivers damaged between 50 and 70 percent of crops (Barnett 2007).

ne of the consequences of natural disasters and sea-level rise is the negative impact on plankton productivity, which results in a decline in fish productivity. For the Pacific Islands, this situation can have drastic consequences because fish exports constitute most of the exports from many countries, such as the Federated States of Micronesia (95 percent), Palau (73 percent), and Samoa (61 percent) (Barnett 2007). Since 1976, more frequent and intense El Niño events and fewer Las Niñas have reduced Pacific tuna catch volumes (Reti 2008), threatening the tuna industry, which is responsible for 11 percent of the combined GDP of all Pacific Island countries and half of the value of all exports from the region (Gillett et al. 2001).

2.2.4 Impacts of climate change on food prices and net trade -- combined IFPRI crop modeling, neural network and IMPACT results

Per capita demand for cereals in Asia declines across the board in the baseline scenario, while per capita consumption of meat increases (Table 2.7). Climate change results in a further decline in 2050 compared to the baseline, with increased prices in the climate change scenario causing per capita consumption to decrease for both cereals and meats. The population increases in South Asia causes total cereal demand to strongly increase there while Central and Southeast Asia only see modest increases in total demand (Figure 2.17). East Asia, in contrast, will see a slight decline in total cereal demand. Total meat demand will be increasing strongly in Asia except in the central region (Figure 2.18).

The price impacts of climate change on major grains are shown in Figures 2.19-223 and in Figures 2.23-2.25 for major livestock products. Overall, climate will have a strong effect on cereal prices and a moderate impact on livestock prices; both cereal and meat prices will increase with climate change as compared to the no climate change scenario. Price increases for rice, wheat, and soybeans due to climate change are on the order of 20 to 70 percent by 2050 compared to the no climate change baseline, while maize, a is projected to be hit especially hard by climate change with prices more than doubling in 2050 compared to the baseline. Livestock prices are affected much more modestly—roughly a 10 percent increase—by climate change in this modeling due to the fact that they are only receiving indirect impacts through changing prices for other crops (both feed and complements/substitutes). Other modeled crops see similar

Page 35: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

35

price changes both in terms of direction and magnitude with only a few seeing impacts as severe as those for maize.

Net trade in meats and cereals in Asia will also see strong adjustments due to climate change (Figures 2.26-27). Under the baseline, Central and Southeast Asia will remain modest exporters and importers, respectively, of both meat and cereal out to 2050, while both South and East Asia will become strong importers of both. Climate change is projected to further increase net cereal imports into East Asia while South Asia would see a decreased dependence on imports. Climate change will slightly reduce meat imports into East and Southeast Asia and India. Other countries in South Asia (Bangladesh, Nepal, Afghanistan, and Pakistan), on the other hand, will need to import more meat.

2.3 Food accessibility

People’s ability to gain access to enough food is a function of market mechanisms—such as cost of food and income-generating capacity—and nonmarket mechanisms—including production for own consumption, food preparation, and allocation practices within the household other than public or charitable distribution schemes (FAO 2008b). In this section, the impacts of climate change on food prices, farm income level, and labor productivity are assessed to investigate the market dimension of impacts on food access.

Nonmarket mechanisms are represented by impacts on gender and land tenure, two important factors that influence food accessibility. Subsistence production and food preparation fall heavily under women’s responsibilities, and in many parts of the world, when a family faces food scarcity, food is preferentially allocated to able-bodied male adults, who are assumed to need it the most to work and maintain the family (FAO 2008b). This example is only a small fraction of how women might be affected by climate change. Gender permeates all aspects of food security and should always be considered in adaptation strategies.

2.3.1 Market dimension

There is a consensus among modeling studies that overall output potential in developing countries will decrease, with consequent increases in food prices, even though studies disagree over the magnitude of climate change impacts (Parry 2007). Climate change is likely to cause food prices to increase moderately until 2050 and at a faster rate after that, when temperatures are expected to increase more substantially (Schmidhuber and Tubiello 2007). By 2080, different scenarios project world food prices to increase from around 7 to 20 percent with CO2 fertilization and from around 40 to more than 350 percent without (Figure 2.30. In developing countries, cereal prices are projected to increase, even if farmer adaptation is taken into account (Rosenzweig et al. 2001).

Climate change might also indirectly contribute to increased food prices through biofuel production, which is one of the mitigation options available to reduce greenhouse gas emissions. Between 2000 and 2007, increased biofuel demand is estimated to have accounted for 30 percent of the increase in global grain prices—39 percent of the increase in global prices of maize, 21 percent of the increase in rice prices, and 22 percent of the increase in wheat prices (Rosegrant 2008).

Page 36: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

36

Increases in food prices directly affect household food security by affecting daily food intake. Higher food prices might result in calorie deficits and micronutrient malnutrition, because low-income people are likely to replace micronutrient-rich foods (animal products, fruits, and vegetables) with less expensive staples (Cohen et al. 2008).

Few studies project the impacts of climate change on farm income in the Asia and Pacific countries. Studies in India found decreases in agricultural income of 3–6 percent (Sanghi, Mendelsohn, and Dinar 1998) and 7–9 percent (Kumar and Parikh 1998) when temperature increased by 2.0ºC. With a more substantial temperature increase of 3.5ºC, agricultural income declined by 3–8 percent in one study (Sanghi, Mendelsohn and Dinar 1998) and 20–26 percent in another (Kumar and Parikh 1998).

Recent Ricardian studies show that in Sri Lanka impacts on agriculture and smallholder revenues will be highly dependent on region and precipitation levels (Kurukulasuriya and Ajwad 2007; Seo, Mendelsohn, and Munasinghe 2005). Overall temperature increases are predicted to be harmful and rainfall increases beneficial. Simulations by Kurukulasuriya and Ajwad (2007) show that at the national level and assuming no CO2 effects, smallholder revenues are expected to increase by 30 percent, assuming a very large increase in precipitation and very slight warming (PCM scenario). Medium warming and small increases in precipitation might lead to gains of 11 percent (CCSR scenario) or 6 percent (CSIRO). On the other hand, large reductions in precipitation and a 3.3oC increase in temperature are likely to reduce net revenues per hectare by 23 percent (HAD3). Impacts on agriculture vary considerably depending on the region within Sri Lanka. In dry regions—the Northern and Eastern Provinces—impacts on agriculture are expected to be significantly negative, whereas in cooler regions—the central highlands—impacts might be zero or positive (Kurukulasuriya and Ajwad 2007; Seo, Mendelsohn, and Munasinghe 2005).

According to Seo, Mendelsohn, and Munasinghe (2005), the national impacts of climate change on net revenue considering the four most important crops to Sri Lanka (paddy, coconut, rubber, and tea) vary from –20 percent to +72 percent depending on GCM scenario (CCSR, CGCM, CSIRO, HAD3, PCM). Kurukulasuriya and Ajwad (2007) present a more pessimistic range of results for smallholder farming, from losses of 67 percent to gains of more than 40 percent (CCSR, CSIRO, Had3, PCM). Neither study considered the beneficial effects of CO2 fertilization (Seo, Mendelsohn, and Munasinghe 2005; Kurukulasuriya and Ajwad 2007).

Taking into account past impacts of climate variability, it is likely that more frequent and severe extreme events will have the greatest impact on smallholder farmers planting rainfed crops, who constitute the majority of farmers in South Asia and have no or very limited adaptive capacity (Kelkar and Bhadwal 2007). In a drought year, one study shows that in three states of eastern India (Chattisgarh, Jharkhand, and Orissa), farm households’ losses of agricultural income range from 43 to 79 percent (data from 1970–2002) (Pandey et al. 2007). Coping mechanisms, including working in nonagricultural sectors, selling asset, and borrowing, usually generate additional revenue. The recovery of total household income loss, however, is estimated to be only 6–13 percent. Overall income losses in the three states were estimated to be in the range of 24 to 58 percent during drought years, compared with normal years. As a result, during drought years the incidence of poverty in those states is substantially higher than in normal years, with a percentage-point increase of 33 percent in Chattisgarh, 12 percent in Jharkhand, and 16 percent in Orissa.

Page 37: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

37

Floods have also been shown to affect rural incomes. In regions severely inundated in Bangladesh, wages decline sharply (Banerjee 2007). For instance, in the monsoon of 1998, wages declined by 30 percent in the district of Comilla, 19 percent in Dhaka, 9 percent in Khulna, and 9 to 19 percent in Rangpur. Wages decreases were directly associated with decreases in crop yield. Furthermore, wages stayed below the normal nonflood averages in the postflood months. On the other hand, in the months that the districts were either marginally inundated or received heavy rainfall but no flooding, agricultural wages increased above the normal nonflood averages (July). The wage increases were associated with higher agricultural yields as a result of improved soil conditions generated by the rainfall.

Under the A2 scenario, the greatest labor productivity losses as a result of climate change will take place in Southeast Asia, Andean and Central America, and the Caribbean as a result of increased outdoor and indoor heat loads that may impair workers’ health and productivity (Kjellstrom et al. 2008). By 2080, losses in work ability of population-based labor, compared with a scenario without climate change, are expected to be 17 percent in Southeast Asia, 7.5 percent in South Asia, and 6.3 percent in East Asia.

2.3.2 Nonmarket dimension

Gender

Climate change and variability are not gender neutral. Among the most adversely impacted groups are rural women in developing countries who depend on constant interaction with the environment for subsistence (Dankelman 2002). Often women are responsible for the basic necessities of the households such as food, fuel, and water. For that, they rely heavily on natural resources (Sachs 2009). Women are also especially vulnerable to natural disasters and climate change because they constitute 70 percent of the population living below the poverty line (Dankelman et al. 2008; Mitchell, Tanner, and Lussier 2007; Dankelman 2002).

Women are much more likely to die in climate events such as flooding and hurricanes than men. In the 1991 cyclone and flood disasters in Bangladesh, for instance, the death rate for women aged 20–44 was 71 per 1,000, whereas the death rate for men was 15 per 1,000 (Dankelman et al. 2008). In the 2004 tsunami in Southeast Asia, the death rate for women averaged three to four times that for men (Bannon and Naraghi-Anderlini 2009). The more powerful the natural disaster, the stronger the impact on female life expectancy relative to that on male life expectancy. Furthermore, the higher the women’s socioeconomic status, the weaker the effect of the disaster on the gender gap in life expectancy (Neumayer and Plumper 2007).

Access to resources. One of the reasons that women are more vulnerable than men to climate events is that they have less access to resources. Countries with a lower female human development index (extracted from the Human Development Index) have considerably more human suffering and killed and affected people, as a result of natural disasters, than countries with high female human development indexes (FHD) (Noy 2009). In the Ganges River basin that connects Nepal, India, and Bangladesh, more intense and severe floods have disproportionately affected women because they have less access to money, land, food, protection from violence, education, and health care (Mitchell, Tanner, and Lussier 2007). With less education and fewer connections to government agencies and powerful people, women are also less likely than men to be able to obtain disaster assistance to rebuild their livelihoods. After floods in Bangladesh, harvest and livestock losses have a disproportionate impact on women, as many of them rely on

Page 38: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

38

food processing, cattle, and chickens for their cash income (Cannon 2002). After natural disasters, women recover more slowly than men from economic losses, and their economic insecurity also increases more than men’s (Peralta 2008). Therefore, women’s socioeconomic status often leaves them more exposed to climate shocks, making it harder for them to restore their livelihoods.

During drought and flood periods, women often need to resort to private lenders for credit, who charge exorbitant interest rates (Mitchell, Tanner, and Lussier 2007). In the Philippines, intense rains and sudden floods and droughts destroyed maize fields and made women unable to repay village moneylenders. Creditors sued the women, and some women consequently went to jail (Peralta 2008). In the province of Pampanga in the Philippines, women farm workers had to prostitute themselves to save their families from starvation, a phenomenon that was called “sex for rice” (Peralta 2008).

Lack of training and education also makes it harder for women to implement adaptation measures. In Bangladesh, India, and Nepal, women consider crop diversification and adapted agricultural practices essential to coping with floods. They feel they lack the technical capacity, however, to adopt agricultural practices (Mitchell, Tanner, and Lussier 2007).

In the Philippines, the national nonprofit organization Amihan notes three ways in which climate change affects women farmers more than men farmers (Peralta 2008). As women own or control fewer resources than men (such as land), they have fewer assets to sell to compensate for crop losses because of floods or droughts. Also, women are more likely to become indebted as a result of climate-induced crop failures because they are much more likely to borrow from informal moneylenders owing to their more limited access to formal credit. Finally, when there are food shortages, women prioritize the food needs of male household members and children over their own (Peralta 2008).

Differences in male and female agricultural labor distribution. Men’s and women’s different roles in agriculture also contribute to women’s vulnerability. In many regions of the developing world, men tend to work in cash crop production, especially highly mechanized ones, while women work with subsistence crops. Women also, however, often work with their husbands in producing cash crops (Sachs 2009). Usually, men are responsible for plowing the fields and driving draught animals—mechanized tasks—whereas women do most of the sowing, weeding, harvesting, and threshing (FAO 2008d; Peralta 2008). Women also are responsible for household food production and small-scale cultivation of cash crops (FAO 2008d). In general, in 65 percent of Asian households, women are responsible for food production (Mitchell, Tanner, and Lussier 2007).

To fulfill their obligations, rural women tend to work longer hours than men, because they are involved in all stages of food production. For instance, in Southeast Asia, 90 percent of labor for rice cultivation is provided by women. In Pakistan, 50 percent of rural women cultivate and harvest wheat. In the Philippines, 70 percent of paid and unpaid agricultural labor in rice and corn production is performed by women (Peralta 2008). In villages in Bangladesh, while women work almost 12 hours a day, men work from 8 to 10 hours. In Africa and Asia, women work 13 hours more than men each week (FAO 2008d). Furthermore, subsistence crops are often the sole sources of food and income of rural women in many regions (Parikh 2007). Therefore, impacts of climate change and climate variability such as reduced crop yields, livestock death, and

Page 39: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

39

production losses are likely to particularly affect rural women, increasing their workload, budgetary problems, and health problems, among many other impacts (Dankelman et al. 2008).

Despite the fact that women engage in a high share of agricultural activities, they have little decision-making power. Men are also more likely to be vested with water rights and own their private equipment, which makes it significantly harder for women to gain access to irrigation systems for vegetable gardens and subsistence crops (Lambrou and Piana 2006; Sachs 2009).

Climate change is also likely to increase the burden on women in regions experiencing water stress and forest depletion (Sachs 2009). Women are mainly responsible for collecting, carrying, purifying, and supplying water (Brody, Demetriades, and Esplen 2008; Parikh 2007). In many regions of the world, women spend up to five hours a day collecting water and fuelwood and up to four hours preparing food (FAO 2008d). In contrast to rural men, who are almost never responsible for water collection in rural areas, women spend long hours carrying heavy containers and as a result suffer from acute physical problems and have limited time for involvement in other activities, such as education, income generation, recreation, and political activities (Khosla and Pearl 2003). In Nepal, girls spend around five hours per week collecting water. In rural India and Africa, 30 percent of women’s daily energy intake is spent carrying water (Ray 2007). Climate change might force women to walk longer distances to collect water. In southwest Bangladesh, increased salinization of drinking water sources is forcing women to travel longer distances on foot, up to 10 kilometers every day, in search of water (Dankelman et al. 2008).

Women’s increased workload and more time collecting water and fuel mean less time available for food production and preparation, which will likely affect household food security and nutrition (Sachs 2009). Floods and droughts make activities typically carried out by women, such as growing and preparing food, collecting water, gathering fuel, and caring for the ill, still more grueling and time-consuming (Peralta 2008).

Cultural and social gender roles. Existing cultural practices or gender inequalities might also intensify the impacts of climate change on women. In most societies, women have fewer ownership rights than men, and often the use of rights by women has to be mediated by their relationships with men (Sachs 2009). In India, Nepal, and Thailand, fewer than 10 percent of women farmers own land (FAO 2008d). After a natural disaster, social practices that curtail women’s rights to land usually grow more severe (Bannon and Naraghi-Anderlini 2009). The same is likely to happen under climate change, which might reduce land quality in many regions of the developing world, increasing the competition for good land.

Examples from past natural disasters show how cultural factors make women more vulnerable to climate events. For instance, during the 1991 cyclone in Bangladesh, a significant number of women died because cultural norms prevented them from leaving the house without male accompaniment. It is estimated that 90 percent of the victims of the 1991 cyclone in Bangladesh were women and children (Cannon 2002). In the tsunami in Sri Lanka, many more women died than men because they did not know how to swim or climb trees (Sachs 2009). After a food crisis, it is also common that women and girls reduce their intake in favor of males, increasing the incidence of malnutrition among women (although famine mortality rates are higher among men because of their higher nutritional requirements) (Bannon and Naraghi-Anderlini 2009).

Page 40: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

40

Access to information. In general, poor rural women have less access to knowledge, information, and technologies than men and richer women. Rural women often face more obstacles in gaining access to agricultural services and information and in participating in organizations (Sachs 2009). Frequently, extension workers in agriculture speak only to men, expecting that men will pass the information to their wives. In many cases, even when women attend community-level meetings, they often do not express their opinions or are not heard. Women on water management boards in India choose not to attend the board meetings and send male relatives instead (Sachs 2009). At the same time, elite women’s demands sometimes override those of poor women, depriving them of rights to resources such as water and fuel.

With less access to extension and markets for seeds and inputs, women farmers are less likely to have access to information about changing weather patterns or new varieties and farming practices that would enable them to adapt to climate change. As a result, they probably receive less warning or information about natural hazards and risks (Bannon and Naraghi-Anderlini 2009).

Leadership roles for women. Women are often the first to suggest and implement strategies to cope with climate change and variability. Their knowledge and experience are valuable inputs that should be used in adaptation strategies. Women’s voices are largely absent, however, from climate change discussions and policies.Women should be perceived not only as victims, but also as positive agents of change in relation to both mitigation and adaptation measures (WEDO 2007; Peralta 2008). Rural women’s responsibilities in food production and preparation and water and fuel collection make them better at developing important strategies for coping with climate change and variability (WEDO 2007; Mitchell, Tanner, and Lussier 2007). Women have broad knowledge and experiences regarding their environment that should be considered in formulating adaptation strategies (Carvajal-Escobar, Quintero-Angel, and Garcia-Vargas 2008).

When they have the means, women are often the first to implement local adaptation strategies. In Nepal, women started growing off-season vegetables and bananas because those crops suffer less during floods and droughts than paddy (Mitchell, Tanner, and Lussier 2007). During a drought in the Federated States of Micronesia, women took the initiative to dig a new well with drinkable freshwater (WEDO 2007). In the Philippines, Amihan members have responded to increases in heavy rains and high input prices by cultivating a traditional, indigenous variety of rice. The new variety does not require massive doses of chemical fertilizers and pesticides and is more resistant to pests than commercial varieties (Peralta 2008).

In fact, it has been observed that communities cope better with natural disasters when women play a leadership role in early warning systems and reconstruction (WEDO 2007). In many cases, women lead communities in conserving natural resources and adapting crops to changing soil and climatic conditions. In the Philippines, often it is women who lead their communities and households in developing agricultural coping strategies, mixed cropping and crop diversification, water harvesting and irrigation, and a growing reliance on wild fruits and forest products (Peralta 2008). Women also tend to share information related to the community’s well-being, choose less-polluting alternative sources of energy, and adapt more easily to environmental changes (WEDO 2007). A study in South Asia shows that women have increasingly shared practices such as promoting alternative energy–related technologies (such as solar, biogas, and improved cooking stoves) (Mitchell, Tanner, and Lussier 2007).

Page 41: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

41

Including women in climate change policies and programs. Currently, there is little evidence that bilateral and multilateral programs fund women’s adaptation activities (Mitchell, Tanner, and Lussier 2007). Neither the Kyoto Protocol nor the UNFCCC mention gender issues (Jamting 2008).

At the implementation level, lack of resources and economic and institutional barriers constrain women from taking more adaptation measures (Carvajal-Escobar, Quintero-Angel, and Garcia-Vargas 2008). In Bangladesh, the Grameen Bank realized that women’s access to irrigation water was ineffective without access to land, credit, seeds, and fertilizer. When all these resources were provided to women, their income from irrigation activities substantially increased (WEDO 2007).

Many policies, legislation, and implementation practices in developing countries need to be changed to adequately address rural women’s adaptation needs. Community and production functions that rural women perform should be used as sources of knowledge in adaptation and mitigation activities (Carvajal-Escobar, Quintero-Angel, and Garcia-Vargas 2008). Furthermore, it is essential that adaptation strategies consider gender issues, because adaptation measures may increase the workload of already overburdened women (Carvajal-Escobar, Quintero-Angel, and Garcia-Vargas 2008).

Adaptation funds should also play a key role in promoting women’s rights by prioritizing the needs of poor women (Peralta 2008; Mitchell, Tanner, and Lussier 2007). For adaptation funding to be efficient, legislation and policies need to protect women’s rights, such as access to knowledge and skills, land ownership, equal participation in decision making, and access to services such as agricultural extension.

Land tenure

It is likely that climate change and variability will change land suitability for crop and livestock production with implications for the land tenure security of poor people, although the linkages between climate change and land tenure are multiple, complex, and indirect (Quan and Dyer 2008).

Sea-level rise in low-lying coastal regions and deltaic floodplains is likely to be a major cause of displacement. Land will become scarcer if low-lying coastal areas disappear temporarily or permanently as a result of constant flooding (FAO 2008c). Because widespread coastal land protection and reclamation might be impossible, the main challenge will be to resettle displaced communities and provide them with alternative livelihoods (Quan and Dyer 2008). In Bangladesh, where land distribution has become more unequal in recent years and likely to become still more so with sea-level rise (FAO 2008c), climate change might disproportionately affect poor people.

Indigenous people are particularly vulnerable to the impacts of climate change because governments often fail to recognize their right to land. Climate change might affect their livelihoods in polar areas, where climate change is melting ice sheets and affecting fish and wildlife populations; mountain environments, where glacial lake outbursts are becoming more frequent, affecting water supply; and tropical forests, which are facing drying trends (Quan and Dyer 2008). Nonexistent or confusing tenure rights may cause those populations to lose their land or be forced to move and not be able to claim income from land-based mitigation or adaptation funding.

Page 42: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

42

Climate change might also substantially affect pastoralists, because climate extremes such as floods and droughts might impair their mobility. Furthermore, they may have to travel long distances as a result of more fragile grazing and scarce water resources. Therefore, adequate forms of tenure and institutional arrangements to guarantee pastoralists’ flexibility need to be put in place (Quan and Dyer 2008).

Indigenous people may also risk losing their land as a result of carbon emissions mitigation projects. Clean Development Mechanism (CDM) projects have not prioritized secure land rights for indigenous people nor have they sought free, prior, and informed consent from these groups before avoided-deforestation or afforestation projects take place (Quan and Dyer 2008). The expansion of biofuel production, another climate change mitigation effort, might have serious implications for land tenure rights of poor farmers. Although small-scale farmers could benefit from biofuel production through increased yields and incomes, those benefits will materialize only if there is secure land tenure. Increases in large commercial biofuel production might result in small farmers’ losing their access to land, which would have a direct effect on food security. In Indonesia, palm oil cultivation has widely hindered land access for local groups (FAO 2008c).

The growing demand for liquid biofuels might also increase the conversion of marginal lands to biofuel production. In India, the government is implementing a national biodiesel program that expects to bring around 400,000 hectares of marginal lands under cultivation. Marginal lands in India, however, supply essential subsistence functions, such as food, fodder, and fuelwood, to the most vulnerable populations (Rossi and Lambrou 2008). Women in particular are likely to be displaced from marginal lands on which they depend for subsistence (Rossi and Lambrou 2008; FAO 2008c).

Finally, lack of property rights makes farmers reluctant to invest in measures to conserve land, as they cannot secure future rights (FAO 1994). Insecure land tenure reduces incentives to improve practices to cope with environmental degradation (Sachs 2009), which intensifies the adverse impacts of climate change and variability on crop production. Unsustainable land practices increase land degradation, which has been associated with global warming.

2.4 Food utilization

Food insecurity is usually associated with malnutrition, but water- and vector-borne diseases also affect people’s physiological capacity to obtain necessary nutrients from the foods they consume (FAO 2008b). Therefore, disease affects how people utilize food, contributing to increased malnourishment rates. This section uses examples from the Asia and Pacific region to analyze the interactions between climate and disease and how climate change will affect patterns of disease. It also highlights some of the direct impacts of climate change on malnutrition rates and the vulnerability of malnourished populations to illness.

2.4.1 Global burden of disease

Climate-sensitive diseases have caused many deaths across the world, particularly in developing countries. Every year, about 800,000 people die from diseases attributed to urban air pollution, 1.8 million people die from diarrhea and poor hygiene, and 3.5 million from malnutrition (WHO 2008). Estimates show that warming and precipitation trends due to anthropogenic climate

Page 43: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

43

change that occurred in the period 1970–2000 claimed 150,000 lives and cost 5 million disability-adjusted life years (DALYs)6 per year (Campbell-Lendrum et al. 2005).

The possible impacts of climate change, including from increased extreme events, on health include increased risk of malnutrition due to decreases in staple food production in poor regions; increased risk of water-borne disease in some regions due to more variable precipitation patterns; alteration of the geographic range of vector-borne diseases; deaths and illness due to thermal extremes; increased burden of diarrheal diseases; increased number of people at risk of dengue; ill health due to social dislocation and migration; and injuries from floods and storms (Confalonieri et al. 2007; WHO 2007; Woodward, Hales, and Weinstein 1998; WHO 2008).

Extreme events can increase the incidence of vector-borne diseases because mosquitoes generally need access to stagnant water to breed. Heavy rainfall and floods can create breeding sites (or on the other hand, wash away those sites), and droughts can increase breeding sites in stagnant water in rivers (Hales, Edwards, and Kovats 2003). In the central highland district of Jayawijaya in Indonesia, after a drought in August 1997, an outbreak of malaria occurred in remote areas of steep mountainous terrain inhabited by shifting agriculturist populations. The outbreak was attributed to various transient pools of standing water along zones of step gradient streams associated with fast-flowing water (Hales, Edwards, and Kovats 2003).

Several studies have positively associated climate-related diseases with El Niño and La Niña–related events. In Bangladesh, changes in the intensity and frequency of the El Niño cycle in past decades have been accompanied by that cycle’s increased role in the temporal dynamics of cholera outbreaks (McMichael, Woodruff, and Hales 2006). Catastrophic floods in Bangladesh have also been associated with diarrhea, respiratory infections, and severely malnourished children (Hales, Edwards, and Kovats 2003). On the islands of Fiji, increases in diarrhea were associated with extreme rainfall events (Singh et al. 2001).

Because mosquitoes breed in standing water, rainfall plays an important role in malaria epidemiology (van Lieshout et al. 2004). In the Indian subcontinent, before the introduction of residual insecticides, malaria epidemics were associated with excessive and failing monsoon rains related to El Niño. From 1868 to 1943, in arid Punjab, excessive rainfall facilitated breeding and increased the lifespan of the mosquito vector, whereas in the wet part of Ceylon, failing monsoon rains caused rivers to pool (Bouma and van der Kaay 1996).

Dengue epidemics have been shown to have significant associations with El Niño in Thailand and La Niña in the South Pacific Islands (Cazelles et al. 2005; Woodward, Hales, and Weinstein 1998). A study that investigated the relationship between reported incidences of dengue fever and El Niño Southern Oscillation (ENSO) in 14 island nations of the South Pacific shows that in the majority of the islands, there were positive correlations between SOI (Southern Oscillation Index) and dengue fever cases (positive SOI is associated with La Niña conditions, when much of the Central Pacific tends to be wetter and warmer than usual) (Hales et al. 1999). In addition to increasing disease transmission, extreme events can also damage or destroy public

6 DALYs correspond to the sum of years of potential life lost due to premature mortality and the years of

productive life lost due to disability.

Page 44: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

44

health infrastructure, especially in small island developing states and other low-lying countries (WHO 2008).

As mentioned, high exposure to diarrhea and other infectious diseases, as a result of floods and droughts, lowers people’s capacity to utilize food effectively, leading to malnutrition (Cohen et al. 2008). When people suffer from poor health and malnutrition, they become too weak to tend their land and the cost of treatment strains their savings (Magrath et al. 2008). Epidemics such as malaria affect the amount of labor available for agricultural activities (Easterling et al. 2007). Therefore, diseases, malnutrition, and the socioeconomic effects of extreme events all reinforce each other and reduce the ability to work. As a result, families’ incomes cannot provide for basic needs, and hunger is a direct consequence.

Mean temperature increases have also been associated with disease outbreaks. From 1986 to 1994 in 18 Pacific Island countries, a positive association was found between annual average temperature and the rate of diarrhea reports and a negative association between water availability and diarrhea rates (Singh et al. 2001). In Fiji in particular, in the period 1978–1998, there were positive associations between diarrhea reports and temperature. An estimated 3 percent increase in diarrhea reports was associated with a 1oC increase in temperature in the previous month (Singh et al. 2001). According to a simulation study in China, climate change is likely to widen the area where schistosomiasis transmission occurs (Zhou et al. 2008). The study projects that global warming will push schistosomiasis transmission into currently non-endemic areas in the north of China, with an additional risk area of 783,883 km2 by 2050. The total area susceptible to schistosomiasis transmission will correspond to 8.1 percent of the surface area of China.

Vector-borne disease transmission is also sensitive to temperature fluctuations. Temperature increases reduce the time required for vector populations to breed and decrease the incubation period of the pathogen (malaria parasite, dengue, or yellow fever virus), which makes vectors become infectious quickly (Hales, Edwards, and Kovats 2003). In the Northwest Frontier Province in Pakistan, increases in mean temperature, rainfall, and humidity were positively correlated with malaria outbreaks of the type Plasmodium falciparum (Bouma, Dye, and van der Kaay 1996). Very high temperature and dry conditions, however, can also reduce the lifetime of mosquitoes (Hales, Edwards, and Kovats 2003).

Box 2.3. Projected impacts of climate change on malaria and dengue transmission

Malaria is strongly influenced by climate conditions (WHO 2008). Climate change might affect malaria’s distributional and seasonal transmission by

increasing its distribution in new areas where it is currently limited by low temperature;

decreasing its distribution where it becomes too dry for mosquitoes to be sufficiently abundant for transmission;

increasing or decreasing the months of transmission in areas with ‘‘stable’’ malaria; and

increasing the risk of localized outbreaks in areas where the disease is eradicated but vectors are still present, such as in Europe or the United States (van Lieshout et al. 2004, 88).

Climate simulations (HadCM3 climate scenarios) project that populations at risk of malaria transmission might increase in Afghanistan, China, and Pakistan by 2080 (impacts of climate variability were not addressed). East Africa and areas in South America are also expected to see an increase in at-risk populations owing to climate change (van Lieshout et al. 2004). To make these estimations, current malaria control status was used as an indicator of adaptive capacity. According to the study, the largest

Page 45: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

45

current population at risk is in Asia, although the burden of disease is concentrated in Sub-Saharan Africa (van Lieshout et al. 2004).

In another study, the area in the world subject to malaria transmission is expected to diminish by 2050. Expansions are predicted in Turkmenistan and Uzbekistan, however, and westward in China (Rogers and Randolph 2000).

Dengue distribution is also highly dependent on climate. In 1990, 1.5 billion people—or almost 30 percent of the world’s population—lived in areas where the estimated risk of dengue transmission was greater than 50 percent (Hales et al. 2002). The same study shows that by 2085, under climate change, 50–60 percent of the projected global population (5–6 billion people) would be at risk of dengue transmission, compared with 35 percent (3–5 billion people) under reference scenarios without climate change. In Asia, at-risk populations are expected to increase the most in the Indian subcontinent, Southeast Asia, and the Pacific Islands—regions that have many areas where dengue is already endemic (Figure 2.31). The onset of the Asian monsoon around June and consequent rainfall facilitate mosquito survival, which is shown by substantial increases in mosquito density in April–July in India and Southeast Asia (Hopp and Foley 2003).

The magnitude of increases in infectious disease burdens due to climate change, however, will depend on several social, economic, and environmental factors. Therefore, a complete assessment of the effects of climate change on malaria and dengue distribution needs to identify not only the areas where climate change might increase transmission, but also those areas where adaptive capacity is low (van Lieshout et al. 2004). Early warning systems and other preventive measures can offer substantial protection against climate-related diseases (Patz et al. 2005).

2.4.2 Overall health impacts

In northern latitudes, climate change might bring benefits such as greater local food production and reduced cold-related mortality and morbidity in winter, but developing countries with already compromised health prospects and declining food yields will see an increase in disease and malnutrition rates (WHO 2008; Confalonieri et al. 2007; WHO 2007; Tschakert 2007; Campbell-Lendrum et al. 2005). The beneficial health impacts of climate change in temperate countries are likely to be greatly outweighed by negative impacts in low-latitude countries, which will contribute to widening the inequality gap between those two groups of countries (WHO 2008).

The gap will depend on countries’ abilities to cope with climate events. For instance, despite the fact that extreme events such as floods can still affect developed countries, they are associated with less physical and disease risks as a result of well-maintained flood control, sanitation infrastructure, and public health measures (Hales, Edwards, and Kovats 2003).

2.4.3 Health impacts on vulnerable people

Climate change is likely to exacerbate current vulnerabilities in countries with poor health systems, preexisting water stress, high population growth, and food shortages (WHO 2008). Many factors will shape vulnerability to climate change, but the effects will be most felt by populations that are already marginal and have low ability to adapt (Woodward, Hales, and Weinstein 1998). High vulnerability can be found in countries such as Papua New Guinea, which has high rates of poverty, low economic development, and rudimentary health and social services (Woodward, Hales, and Weinstein 1998).

Page 46: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

46

Furthermore, other adverse impacts of climate change and variability, such as food and water shortages and increased conflicts, will increase susceptibility to disease, especially in places where vulnerable populations have low adaptive capacity. In the example of Jayawijaya, food and water shortages not only increased the population’s susceptibility to malaria, making the outbreak more intense, but also led to increase demographic movements. People returning to the highlands were thus exposed to high-risk malaria endemic to the lowlands (Hales, Edwards, and Kovats 2003). In Bangladesh, a study found a positive association between drought, lack of food, and increased risk of mortality from diarrhea (Confalonieri et al. 2007).

Most important, susceptibility to climate-related diseases such as malaria increases in populations with poor nutrition and high poverty rates (Woodward, Hales, and Weinstein 1998). There is a striking correlation between malaria and poverty, and malaria-endemic countries also have lower rates of economic growth (Sachs and Malaney 2002). In Bangladesh, lost income due to iron deficiency is estimated to reduce GDP by as much as 8 percent. Assessments in Bangladesh, Guatemala, and Viet Nam show that there is a higher incidence of illness in poor households than richer ones (Cohen et al. 2008).

Poor women and children are likely to bear most of the disease burden of climate change (WHO 2008). Diseases such as diarrhea, vector-borne diseases, and infections due to undernutrition are more severe in children living in poverty (WHO 2008). Almost 90 percent of the burden of malaria and diarrhea and almost the entire burden of diseases associated with undernutrition are borne by children aged five years or less, mostly in developing countries (WHO 2008). Because women are responsible for washing and water collection in many countries, they are particularly vulnerable to water-associated diseases (Poverty-Environment Partnership 2003). The elderly and people with infirmities or preexisting medical conditions are also likely to be among the most severely affected by the health effects of climate change (WHO 2008).

Therefore, the magnitude of climate change’s impacts on health will depend on the current health status of populations in Asia and Pacific countries. About 70 percent of people in Southeast Asia and 95 percent in South Asia are at risk of low zinc intake, which is much higher than the figure for developing countries overall (60 percent). Inadequate dietary zinc can have serious consequences, including stunting (lower than expected height for one’s age), which can lead to higher rates of illness and death, reduced cognitive ability and school performance in children, and lower productivity and lifetime earnings for adults (Cohen et al. 2008).

As a result of chronic undernutrition, 178 million children under the age of five suffer from stunting, and 85 percent of them live in 20 countries in Africa and the Asia-Pacific region. Except for Yemen, the Asia and Pacific countries are all in South Asia (Afghanistan, Bangladesh, India, Nepal, and Pakistan) and Southeast Asia (Indonesia, Myanmar, the Philippines, and Viet Nam). In addition, more than 75 percent of pregnant women in South and Southeast Asia have anemia, which poses a serious risk to mothers during childbirth and affects the child’s health as well.

In South and Central Asia, 27 percent of babies are born with low birth weights (less than 2.5 kilograms), which is higher than the overall average for developing countries (16 percent). Between 1990 and 2005, child malnutrition fell in all developing countries, except in Sub-Saharan Africa. In South and Central Asia, however, the number of malnourished children and

Page 47: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

47

the prevalence of child malnutrition (67 million and 37 percent, respectively) exceeded the levels for of all Africa (34.5 million and 25 percent) (Cohen et al. 2008).

A recent study estimates that 2.2 million deaths and 21 percent of DALYs for children young than five years are caused by stunting, severe wasting, and intrauterine growth restriction every year. Deficiencies of vitamin A and zinc are estimated to be responsible for 1 million deaths and 9 percent of childhood DALYs yearly, worldwide. Iron and iodine deficiencies account for 0.2 percent of global childhood DALYs. Iron deficiency, a risk factor for maternal mortality, was estimated to cause 115,000 deaths and 0.4 percent of DALYs, while suboptimum breastfeeding was estimated to cause 1.4 million child deaths and 44 million DALYs. All those factors together were estimated to be responsible for about 35 percent of child deaths and 11 percent of the total global disease burden (Black et al. 2008). The cost of such diseases is high. In general, difficult pregnancies and illnesses related to malnourishment of mothers and children cost around US$30 billion annually (Cohen et al. 2008).

Some Asian countries have reduced the number of people suffering from undernutrition in recent years. Southeast Asia has significantly decreased undernutrition rates since the 1990s. China has also lowered its rate from 15 percent (1990–92) to 9 percent (2003–05) (Table 2.8). Other countries, however, are still struggling with increasing rates. India reduced its proportion of undernourished people from 24 to 21 percent; however, the total number of undernourished people increased from 206.9 million to 230.5 million. Indonesia (in contrast to the trend in Southeast Asia), Mongolia, Pakistan, Tajikistan, and Uzbekistan also saw an increase in the total number of undernourished people. Uzbekistan has had one of the largest relative increases in the total number and proportion of undernourished people in the total population. In that country, from 1990–92 to 2003–05, the total number of undernourished people increased from 1.0 to 3.6 million, corresponding to an increase from 5 to 14 percent (Table 2.8) (FAO 2008e).

Finally, nonclimatic factors that already have a major influence on disease transmission in developing countries will intensify the impacts of climate change. High population density, forest clearance, irrigation, movement of people, urbanization, resistance to insecticides, resistance to antimalarial drugs, degradation of health infrastructure, conflicts, and political and economic factors are all among the factors that have been associated with disease transmission (Reiter 2001).

2.4.4 Impacts of climate change on calorie availability and malnutrition -- combined IFPRI crop modeling, neural network and IMPACT results

Drawing upon the policy modeling framework of the IMPACT model, we can quantify the effects of climate change on agricultural economies and welfare, and keep track of the evolution of key welfare trends, such as that of malnutrition among children aged 5 years and under7. The basic relationship within the IMPACT model that accounts for the number and prevalence of malnutrition among this vulnerable demographic group, can be summarized as follows:

ln , ,ma utrition f calorie availability clean water female schooling

7 This measure has been used in a number of IFPRI projections (Rosegrant 2001; von Braun 2008) and is defined as the

deviation from a standardized measure of weight-for-age in small children, who are a particularly vulnerable demographic (Smith and Haddad 2000).

Page 48: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

48

where the availability of calories for consumption, itself, is a function of other economic variables coming out of the IMPACT model, such as food prices and food supply8.

This relationship, in essence, embodies at least three of the main components of food security that we’ve discussed in this chapter – namely, availability, access and utilization. The availability component is addressed by the production of food, and its availability to consumers on the agricultural market – further translated into the availability of calories for consumption and intake. The dimension of access is addressed by the relationship of food prices to other market-level drivers – and captures the extent to which consumers can be ‘rationed’ out of the market by high price levels, and driven into hunger. The dimension of utilization is only partially covered by the country-level access to clean water that is quantified in the model, given that there are other important determinants that allow the utilization of available nutrients within the body, such that it can manifest itself in better food security outcomes. Given that clean water is an important determinant of health outcomes – through its effect on reducing prevalence of water borne diseases that often lodge themselves in the gastro-intestinal tract – it does enhance the ability of the body to absorb available nutrients, and leads to overall higher levels of productivity and well-being. While we don’t directly model the process of food preparation, we recognize, as well, that clean water is an essential component of good nutrition outcomes, through its enhancement of food quality at the household level, during the cooking and preparation process.

By connecting the model outputs that were described in section 2.1, under climate change, with the malnutrition component of IMPACT, we see that India, by itself, accounts for over half the total malnourished children within the Asian region, with the rest of the South Asian region accounting for 17 percent of Asia’s child malnutrition. This is a sharp contrast to the East Asia region, where just under 10 percent of Asia’s malnourished children live, and with just over 12 percent located in Southeast Asia. Looking at Figure 2.32 , we see that the effects of climate change, induce a varied range of impacts on the Asian region.

The larger numbers in South Asia translate to only a 3 percent increase over the no-climate change case, whereas the impacts in East Asia (which represent less than a third of the impacts in South Asia) translate to nearly a 50 percent increase over the no-climate change case (Figure 2.33). These climate-driven changes in welfare represent an overall setback in the improvement of welfare in the Asia region, as measured through malnutrition levels of the vulnerable under-5 demographic. Figure 2.33 shows these malnutrition impacts of climate change in more detail for the biggest countries in the East and South Asian regions – China and India (respectively). While the magnitude of climate change impact is relatively close for these countries (an increase of 676 thousand malnourished children for China, and 983 thousand for India, Table 2.9), what they represent a much larger ‘bump’ for China, as a percentage of its no-climate change case – being 57 percent higher, compared to a nearly 3percent increase over the no-climate case for India. In the case where there is no climate change envisioned, the large and fast-growing Asian economy of China sees an 88 percent decrease in the level of malnourished children, over the period. In the case of climate change (as embodied in the A2a SRES scenario), the reduction over the 50 years would become just over 80 percent, which is still a considerable improvement. By contrast, India’s level of child malnourishment would decrease by just half that rate – reaching 63percent of the year 2000 levels of malnutrition by 2050, compared to just under

8 A more complete description of the malnutrition module of IMPACT is given in the modeling appendix.

Page 49: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

49

20 percent for China. Much of the reason for these differences lies in the fact that so many more undernourished (and poor) people are located in South Asia, compared to East Asia – making the burden of overall welfare improvement more onerous.

Generally, malnutrition is expected to decline over the next 50 years. This trend is reflected in the poverty and welfare projections of both the World Bank (World Bank 2007) and other groups (Hughes et al 2008). The results for Indonesia are quite similar to those for India, as are the results for Thailand, as well.

In the case of other countries like Malaysia, the Philippines, Sri Lanka and South Korea, they are also able to attain a reduction in their levels of malnutrition of close to half, under a no-climate change situation, with each of those beginning from a wide range of levels. In the case of climate change, the reduction in hunger is offset by roughly 3 percentage points, which represents a wide range of actual numbers not lifted out of malnutrition, depending on the initial levels of malnutrition that are embodied in those regions.

Those countries able to decrease their levels of child malnutrition by roughly a third over the 50 year period, without climate change, include Thailand, Turkmenistan, Myanmar and Viet Nam. While the effect of climate change on most of these countries would offset these improvements by 1 to 2 percentage points, a country like Turkmenistan would be affected to a greater degree, such that it would only achieve a 22 percent reduction, compared to a 35 percent reduction over the 50-year period. A handful of countries in the Asia region are projected to increase their levels of malnutrition to 2050, in contrast to the overall trends in Asia, such as Afghanistan and the part of Southeast Asia embodied by Cambodia and Laos. Afghanistan, in particular, is hit further by the effects of climate change, such that the increase in malnutrition goes up from 12percent to 37percent by 2050, embodying a total increase of 224 thousand children.

These trends illustrate the parts of Asia that warrant particular attention by policy makers, such that consideration of additional interventions can be made – particularly in the face of imminent stresses and shocks to food systems that are likely to occur under climatic change.

In Chapter 5 we will discuss, more precisely, the kind of interventions and investments that can be made to offset the malnutrition effects of climate change that we’ve illustrated here. This will give us greater insight into the avenues that are available to allow policy action to make a difference in the well-being of those who are most vulnerable to the effects of climate change.

2.4.5 Role of health infrastructure

The health impacts of climate change will be felt mainly by populations without access to adequate health infrastructure (Hales, Edwards, and Kovats 2003; van Lieshout et al. 2004). Inadequate infrastructure and ineffective vector and disease surveillance, among other factors, have been associated with the spread of dengue in urban areas of tropical countries in recent decades (Hales, Edwards, and Kovats 2003). On the other hand, malaria transmission declined in Europe because of socioeconomic development, such as modernization of livestock production and farming (van Lieshout et al. 2004). In the Netherlands, the use of quinine in rural medical care and the stabling of cattle away from human habitations were responsible for the eradication of epidemic malaria, while in England, the reduction of malaria transmission was attributed to a combination of social, economic, agricultural, educational, and public health factors (van Lieshout et al. 2004). Therefore, under areas susceptible to the propagation of diseases such as

Page 50: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

50

malaria, disease incidence will be higher in places where the public health structure is not effective (Hales, Edwards, and Kovats 2003).

In regions with poor water supply services and sanitation, the impacts of extreme events are likely to be more severe. Around 90 percent of the burden of diarrhea diseases is attributable to lack of access to safe water and sanitation (WHO 2008). Cholera transmission in many countries, for instance, is associated with poor sanitation (Confalonieri et al. 2007). Floods can induce outbreaks of cholera, typhoid, and diarrhea diseases if the floodwaters and thus water supplies become contaminated with human or animal waste (Hales, Edwards, and Kovats 2003). Droughts can also increase the risk of disease by reducing the water available for washing and sanitation. During water shortages, water use is prioritized for cooking rather than hygiene, a practice that might increase the risk of diseases, such as diarrhea, that stem from fecal contamination and water-washed ones such as trachoma and scabies9 (Hales, Edwards, and Kovats 2003).

Climate change might also indirectly contribute to increased disease transmission through population displacement. More frequent and intense floods and droughts, competition over land, water scarcity, and poor governance might all force population displacement. Such displacements have been associated with increased risks of several health effects (from mental disorders to communicable diseases) and potential conflict (WHO 2008). In refugee populations, infectious disease outbreaks are common as the result of inadequate public health infrastructure, poor water and sanitation, overcrowding, and lack of shelter (Hales, Edwards, and Kovats 2003). Population displacement might also affect health by pushing populations into regions susceptible to infectious diseases.

Furthermore, people in refugee camps depend on food rations that are often insufficient, making them susceptible to malnutrition. In 2005, surveys indicated that half of refugee camps were considered to be in a state of acute malnutrition emergency (Cohen et al. 2008). In refugee camps in Bangladesh and Nepal, lack of B vitamins is a public health problem (Cohen et al. 2008).

2.5 Food system stability

More frequent and severe droughts and floods, as well as conflicts, will potentially affect food supply stability. Droughts and floods are a particular threat to food supply stability, and conflicts destabilize food systems at all levels. Increased instability of supply as a result of climate change is likely to lead to an increase in food emergencies, with which the global food system is ill equipped to cope (FAO 2008b).

2.5.1 Impacts of climate extreme events

In Asia, a higher incidence of climate extreme events is a particular reason for concern, because statistics for 1975–2006 show Asia as the most disaster-afflicted region in the world. Asia accounted for about 89 percent of people affected by disasters worldwide, 57 percent of total fatalities, and 44 percent of total economic damage. In that period, 75 percent of all natural disasters in Asia were hydro-meteorological disasters (Sanker, Nakano, and Shiomi 2007). In

9 Water-washed diseases are caused by lack of proper sanitation and hygiene. If people were able to wash themselves, their

clothes, and their homes, that could protect them from infection.

Page 51: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

51

2006, 21 of the world’s 25 top natural disasters, in terms of number of people affected, occurred in Asia (Table 2.10). Of those 21 disasters, 11 occurred in China. Afghanistan and China were the only countries affected by droughts in 2006 (Noy 2009). In 2007, 9 of the 10 countries with the highest death rates caused by extreme events were in Asia. The most-affected countries were, in order, Bangladesh, India, China, Pakistan, Korea – DPR, United States, Indonesia, Viet Nam, Afghanistan, and Nepal (Harmeling 2008).

In the lower Mekong River, where predictions of more intense precipitation for the rainy season suggest increases in the magnitude and frequency of flooding (Chinvanno et al. 2006), more frequent and severe climate extreme events are already being felt. In the Ninh Thuan province in Viet Nam, the 2004 drought caused an unprecedented reduction in the irrigated area by 5,000 hectares (ha) and affected agricultural land area of 5,185 ha. On the other hand, inundation in 2003 flooded 15,591 ha of agricultural land (of which 9,190 ha were cultivated with paddy) and killed 45,644 animals (Oxfam and Kyoto University 2007). Paddy and shrimp farming in the Mekong Delta in Viet Nam is likely to be heavily affected by inundation (because of increases in heavy rains) and sea-level rise during the rainy season. Climate studies also project a decrease in precipitation during the dry season and increasing salinization problems in a region in which 42 percent of the area was already affected by salinity intrusion by 1995 (ADPC 2003).

In South Asia, countries are also vulnerable to floods. The dramatic consequences of floods can be seen in countries such as Bangladesh, where annual floods inundate 20.5 percent of the country’s area and up to 70 percent during extreme flood events. The increase in the frequency of natural disasters in Bangladesh has led not only to loss of land directly to the sea, but also to deposits of large amounts of sand and salt on agricultural land as a result of river and coastal flooding. These deposits have led to the abandonment of land in some regions (Ansorg and Donelly 2008). Climate change is likely to increase the frequency of floods in Bangladesh and cause crop losses of 1.5 million metric tons over a 20-year period (Mirza 2002).

In recent years, many inhabited islands in the Maldives have experienced severe weather events such as strong winds, flooding, rough seas, and storm surges resulting in severe beach erosion hazards and coastal infrastructure damage. The country was one of the most severely affected by the 2004 Asian tsunami. A population scattered across 198 islands makes disaster management and food distribution extremely difficult. As a result, vulnerable populations depend on home-grown vegetables and fruits (rice, the main staple food, is imported from other South and Southeast Asian countries) (MPND 2007). Therefore, the Maldives is extremely vulnerable to climate change, especially to a higher incidence of climate extreme events.

In the Pacific Islands, cyclones, storms, heat stress, and droughts are already responsible for significant agricultural production losses because of erosion, increased groundwater contamination, and saltwater incursion into estuaries. For instance, in 2005 in the Cook Islands, saltwater entirely inundated the taro plantation areas in Pukapuka as the result of a cyclone. Only after three years could taro be reintroduced to the island (FAO 2008a). Furthermore, the Pacific Island countries are especially vulnerable to climate change, as increases in natural disasters and sea-level rise are likely to make some areas uninhabitable. Vanuatu, Kiribati, the Marshall Islands, Tuvalu, Fiji, the Solomon Islands, and parts of Papua New Guinea are at serious risk, and Tuvalu and Kiribati could be almost completely submerged by the mid-21st century (Tay 2008).

Page 52: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

52

In Viti Levu in Fiji, sugarcane is likely to be adversely affected by climate change because it is particularly sensitive to droughts. If future droughts in Viti Levu are as intense as the 1997/98 drought and as frequent as they were over the period 1983–98 (one drought every four years), projections for the next 25–50 years are dire. Sugar production is expected to reach a normal level of output (4 million metric tons) only 44 percent of the time (7 out of 16 years), whereas output is expected to be half that of a normal year 25 percent of the time (4 years), and three-fourths of normal yearly output 31 percent of the time (5 years) (World Bank 2000). On the other hand, during El Niño years, yam production is expected to remain the same or even increase.

Of all small islands, atoll countries are the most physically vulnerable to sea-level rise as a result of their high ratio of coastline to land area, high population densities, and low level of adaptive capacity (Barnett and Adger 2003). Atoll countries such as Kiribati, the Marshall Islands, and Tuvalu (and the Maldives in South Asia) are particularly at risk. Increases in sea-surface temperature (SST) damage coral reefs, which are crucial for the formation and maintenance of atoll motu10. As a result of coastal erosion and a decrease in coral resilience, atoll countries are likely to see more floods, increases in contamination of freshwater aquifers with saline water from storm surges and seepage, and decreases in agriculture and artisanal fishing productivity (Barnett and Adger 2003). Therefore, climate change is likely to increase food insecurity in atoll countries by decreasing local production of fish and crops and the ability of those countries to pay for food imports as a result of economic contraction (Barnett and Adger 2003).

A recent study reports that droughts have a direct impact on household food consumption (Pandey et al. 2007). During droughts in eastern India, households are unable to maintain the food consumption level of non-drought years and thus experience increased food insecurity. In drought years, the average number of meal per day drops from close to three to close to two, with 10 to 30 percent of surveyed households reducing their frequency of food intake to one meal per day. Most people (60 to 70 percent) also reduce the quantity of food consumed per meal besides consuming more “inferior” food items.

The impacts of droughts on household nutrition, especially of children, are likely to be life-long. Droughts can increase stunting and malnutrition rates because they are likely to reduce dietary diversity and overall food consumption (Cohen et al. 2008). Further research should investigate the long-term implications of stunting in children of the Asia and Pacific region. In a study in rural Zimbabwe, however, an econometric model was used to assess the negative impacts of the 1982–83 drought on children’s height-for-age variable. Results show that the exposure of 12- to 36-month-olds to that drought event resulted in a average loss of stature of 2.3 centimeters and 0.4 grade of schooling, translating into a 7 percent loss in lifetime earnings (Alderman, Hoddinott, and Kinsey 2003).

2.5.2 Potential for conflicts

Declining food production might turn into a food crisis in some regions, depending on countries’ socioeconomic and political characteristics. A country with high economic output, high per capita income, and a low contribution of the agricultural sector to GDP will be better able to rely

10 According to Barnett and Adger (2003), “atolls are rings of coral reefs that enclose a lagoon. Around the rim of the reef

there are islets called motu with a mean height above sea-level of approximately two meters” (p 322).

Page 53: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

53

on agricultural imports to compensate for drops in production as a result of climate change. Economically weaker countries that depend on agricultural exports as a main source of income have less ability to avert crises or avoid subsequent violence as a result of substantial declines in GDP (Schubert et al. 2008). Political stability and governance of countries, the quality of agricultural infrastructure, migration patterns, civil society structures, wealth distribution, and institutional stability are all factors that will interact to determine potential outcomes related to losses of food production due to climate change (Schubert et al. 2008; Preston et al. 2006).

Climate-change related migration might increase or reduce conflicts. Migration away from a region might reduce land competition among market-oriented agricultural producers and subsistence farmers, reduce environmental degradation, and reduce the risk of violent conflict. On the other hand, migration can be the cause of violent confrontation when the presence of migrants increases the pressure on scarce resources (Schubert et al. 2008). Figure 2.34 shows regions that have already suffered from environmental conflicts as a result of biodiversity, fish, water, and land disputes.

Central Asia is especially prone to conflicts as a result of water disputes among countries. Water resources in the region are defined as national or transboundary. In Uzbekistan, for instance, 90 percent of its river flow is formed beyond its boundaries (Hakimov et al. 2007). About 80 percent of the population in Central Asia depends on the water from the Amy Darya and Syr Darya Rivers (Hakimov et al. 2007). Countries in the region depend on the cultivation of water-intensive crops such as cotton and rice. Agriculture accounts for 90 percent of water use, and around 75–100 percent of agricultural land in the region requires irrigation. As a result, soil salinization in the region has increased. Conflicts over water use and distribution have increased between upstream and downstream countries (Schubert et al. 2008), and countries have often accused each other of exceeding their quotas (Crisis Group 2002). The downstream countries (Kazakhstan, Turkmenistan, and Uzbekistan) have rising populations and are heavy water consumers for cotton production, whereas the upstream countries (Kyrgyzstan and Tajikistan) want to use more water for electricity generation and farming (Crisis Group 2002). Furthermore, even if sufficient water exists in Central Asia to meet its population’s needs, water stress might increase as a result of the deterioration of irrigation and sanitation infrastructure and poor administrative water allocation systems (Hakimov et al. 2007).

Climate change might also lead to increased conflicts in South Asia, where countries have serious socioeconomic and political problems and intra- and interstate conflicts. With Pakistan’s capacity for irrigated agriculture reduced and Bangladesh’s floods, migrations toward India might increase and current conflicts may escalate (Schubert et al. 2008). The 5.5 million people living in the Ganges Delta in Bangladesh will be forced to relocate if sea-level rise reaches 45 centimeters. Many might seek to move inland within Bangladesh. Many others, however, might seek to move to India and Pakistan, where previous migration of this kind has already caused violence (Barnett 2003). Among and within countries, decreased water availability during dry seasons will increase competition for water, generating more conflicts (Schubert et al. 2008).

2.6 Conclusion

Climate change is a major challenge at the global level and in the Asia and Pacific region. Decreased agricultural production in most of the region owing to climate change will result in

Page 54: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

54

higher food prices, tending to depress food consumption, especially among the poor. The result will be an increased number of people at risk of hunger. Regions that are already lagging behind in achieving important human well-being outcomes, such as Afghanistan, will likely suffer the most.

To fully understand the impacts of climate change on rural livelihoods, all dimensions of food security need to be addressed. The impacts of climate change on gender, health, conflicts, and other factors will affect food security and therefore should be better investigated. Furthermore, global warming and changes in rainfall patterns need to be analyzed together with the impacts of rises in sea level, increases in climate extremes, water scarcity, land degradation, and other stressors. Needless to say, impacts will be felt differently by countries according to their adaptive capacities and coping abilities.

As a result of uncertain climate predictions and other factors (CO2 fertilization effects, socioeconomic pathways, countries’ adaptive capacity), projections of the impacts of climate change on agriculture are not as precise as desired and depend heavily on scenarios’ assumptions. Nonetheless, projections show that agriculture systems in many vulnerable regions in the Asia and Pacific countries will suffer with climate change. In particular, global assessments predict losses in agricultural potential and yield for South Asia. On the other hand, some regions, mainly in East Asian countries, are expected to benefit from the combined effect of global warming and CO2 fertilization. Further research should be done to assess impacts in Central and Southeast Asia, as well as in the Pacific Islands. There will always be uncertainty in climate change science; however, it is possible that investments in research to better understand its full dimensions will reduce the uncertainty level.

The Asia and Pacific region is familiar with the drastic economic consequences of natural disasters. More intense and frequent climate extreme events may have severe consequences for rural livelihoods. Considering the countries’ current vulnerabilities, more intense and frequent disasters such as floods and droughts are likely to have a negative impact on agricultural GDP and trade, particularly in developing countries. Severe floods and droughts are also known to lead to declines in rural wages because fields are less productive and farmers’ coping mechanisms often fail to compensate for losses of agricultural revenue. Hence, more intense events will only exacerbate farmers’ vulnerability.

Even without climate change, competition for land and water resources is high in many Asian and Pacific countries. Climate change will intensify the struggle for these natural resources, aggravating their management in the region and increasing risks of conflicts. Central and South Asia are particularly prone to conflicts as a result of water and land scarcity.

Low-income and vulnerable populations will feel the effects of climate change and increases in the incidence of natural disasters most strongly. Rural women will be among the most affected groups in the world as a result of their dependency on subsistence crops and limited access to resources and decision making. Adaptation strategies should acknowledge the greater vulnerability of women to climate change. In fact, for adaptation and mitigation measures to be successful, an assessment of poor communities’ current vulnerabilities, needs, gender aspects, and coping abilities is needed.

Page 55: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

55

CHAPTER 3. VULNERABILITY OF ASIA AND PACIFIC COUNTRIES TO CLIMATE CHANGE

Key Messages

Vulnerability to climate change depends not only on exposure to climate events, but also on physical, environmental, socioeconomic, and political factors that influence how sensitive countries will be to a changing climate and how they will be able to cope.

Climate change is likely to increase the vulnerability of poor farmers who already struggle with land degradation in the Asia and Pacific region.

Countries in South Asia, Southeast Asia, and the Pacific Islands are highly vulnerable to sea-level rise, which will increase the risk of floods. Moreover, glaciers in the Himalayas and Central Asia are already melting as a result of global warming with potential short-term benefits and risks but likely adverse long-term impacts on food production and ecosystem health in the dry season.

In regions highly dependent on livestock production, such as Mongolia and Inner Mongolia, China, overgrazing increases vulnerability to climate change.

A combination of indicators reflecting exposure and sensitivity to climate change and adaptive capacity identifies Afghanistan, Bangladesh, Cambodia, India, Lao PDR, Myanmar and Nepal as most vulnerable to climate change—with poor outcomes in all three vulnerability components—revealing South and Southeast Asia as the regions most vulnerable to climate change in the Asia and Pacific region. Data for most Pacific islands are insufficient to construct the same vulnerability indicator.

Countries with significant vulnerability—poor outcomes in two out of three components—include Bhutan, China, Pakistan, Thailand, Timor-Leste, Uzbekistan, and Viet Nam scattered throughout the Asia and Pacific region.

3.1 Factors affecting vulnerability to climate change in the Asia and Pacific region

This chapter reviews the vulnerability of countries in the Asia and Pacific region based on composite indicators reflecting exposure, sensitivity, and adaptive capacity to climate change, following the Intergovernmental Panel on Climate Change (IPCC) (McCarthy et al. 2001). The IPCC’s definition of vulnerability combines information on potential climate impacts and on the socioeconomic capacity to cope and adapt (O'Brien et al. 2007; O'Brien et al. 2004; Fussel 2007).

Vulnerability assessments show that the poorest and most vulnerable countries and populations are the first and most affected by climate extreme events. While mortality risks are clearly lower in countries with developed economies, high mortality rates and vulnerability to natural disasters have been associated with low-income countries, densely populated areas, inefficient governments, lack of accountability, high levels of inequality, and low literacy rates (Stromberg 2007; Kahn 2005). From 1975 to 2006, for instance, almost 80 percent of all deaths from hydro-meteorological disasters occurred in the lowest-income countries and only 4 percent in high-income countries (Table 3.1). Disasters in Bangladesh, China, India, Pakistan, and some

Page 56: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

56

countries in Africa made significant contributions to this trend (Sanker, Nakano, and Shiomi 2007).

In addition, the economies of developing countries and smaller countries are less able to cope with disasters of similar magnitude, than the economies of developed or larger countries. Countries with higher literacy rates, better institutions, higher per capita income, higher degrees of openness to trade, and higher levels of government spending are better able to cope with the initial disaster shock and avoid spillovers into the macroeconomy (Noy 2009). Marginalized social groups in developing countries, including poor women, children, the elderly, and disabled people, have suffered the most from natural disasters. Hence, expected increases in such events as a result of climate change will disproportionately affect those groups because the impact of such disasters depends not only on exposure, but also on people’s levels of vulnerability (Ehrhart et al. 2008). Furthermore, climate change will hit communities in the Asia and Pacific region that already experience high levels of food insecurity. According to the Food and Agriculture Organization of the United Nations, the Asia and Pacific region accounts for 68 percent of the developing world’s population and 64 percent of its undernourished population. Sixteen percent of the region’s population is undernourished (FAO 2006). Thus, countries with higher adaptive capacity are more likely to weather adverse impacts from climate change.

According to Preston et al. (2006), the Asia and Pacific region is exposed to a range of climate conditions and extreme events. The El Niño-Southern Oscillation strongly influences rainfall patterns in the region, bringing periodic drought and extreme sea levels in the southwest Pacific. Furthermore, tropical cyclones and associated high winds, storm surges, and extreme rainfall events are common in the coastal Asia and Pacific area (Preston et al. 2006). Climate change might significantly alter the dynamics of these events, possibly increasing their frequency and intensity in many countries. Low-lying countries, including small islands, will face the highest exposure to sea-level rise, which will increase the risk of floods that might affect millions of people in the Asia and Pacific region.

Vulnerability to climate change will also be higher in countries where agriculture accounts for a large share of GDP and employment, where levels of poverty are high, and where population density is high. These characteristics apply to many countries in South Asia. Finally, land and water degradation, important causes of crop yield decreases, will also make countries more sensitive to a changing climate.

3.2 Results from vulnerability assessments for the Asia and Pacific region

Globally, during 1980–2004, droughts have been the most deadly geophysical and hydro-meteorological events, followed by windstorms and tsunamis, whereas floods have affected the largest number of people (Stromberg 2007). Flood-risk hotspots were identified in South and Southeast Asia; drought-risk hotspots in South Asia (Afghanistan, Pakistan, and parts of India) and Southeast Asia (Indonesia, Myanmar, and Viet Nam); and cyclone-risk hotspots in Bangladesh, parts of India, Viet Nam, and other Southeast Asian countries (Ehrhart et al. 2008). Thus, many countries in the region, particularly in South and Southeast Asia, have areas at risk from more than one climate-related hazard (Figure 3.1).

Different studies using different methodologies have proposed vulnerability indexes to assess countries’ vulnerability to climate change (see summary in Table 3.2). As described here, some countries are considered vulnerable according to several different criteria. Detailed

Page 57: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

57

scenario analysis of vulnerability in Central Asia and the Pacific Islands is left out as a result of the lack of reliable data and comprehensive studies that investigate vulnerability in these regions. An overview of the most vulnerable countries in the Asia and Pacific region is presented here.

Bangladesh

The low-lying coastline, high population density, and an economy highly dependent on agriculture together make Bangladesh one of the most vulnerable countries to sea-level rise and other effects of climate change (Poverty-Environment Partnership 2003). Bangladesh is a cyclone and flood risk hotspot11 (Ehrhart et al. 2008). Most of the country’s elevation does not exceed 10 meters, and a one-meter sea-level rise might well result in the flooding of 16 percent of the country’s land area (Karim, Hussain, and Ahmed 1996). Moreover, even a sea-level rise of 0.30 or 0.75 meter is expected to wreck havoc in the eastern coast of Bangladesh, flooding areas of 5.80 and 11.20 km2, respectively, 95 percent of which is agricultural land (Ali 1999). Box 3.1 presents further detail on sea-level rise projections for the Asia and Pacific region. Finally, Bangladesh will also be affected by glacier melt, which in the long run could further worsen the impacts from sea-level rise because lower dry-season river flows would further draw in saltwater (see Box 3.2 for impacts of glacier melt under global warming).

About 20 percent of Bangladesh’ GDP comes from the agricultural sector, which employs more than half of the total workforce (World Bank 2008). Furthermore, rural density in the country is extremely high with 1,249 people per km2 of arable land (World Bank 2005).

According to Moss, Brenkert, and Malone (2001), even the country’s current sensitivity to climate change is beyond its adaptive capacity, and by 2095 sensitivity is expected to increase even more in two out of three scenarios (Table 3.2). Similarly, according to Yohe et al. (2006), Bangladesh will be extremely to significantly vulnerable to climate change under all scenarios, including under a scenario that combines mitigation and enhanced national adaptive capacity (Yohe et al. 2006).

Nonetheless, the decline in the number of people killed in the tropical cyclone of 1997—less than 200 people killed—compared with a similar storm in 1991 that led to a death toll of 138,000 show that successful adaptation (in this case disaster management involving governmental and nongovernmental organizations) can significantly reduce a country’s vulnerability to climate events (Brooks, Adger, and Kelly 2005). Ehrhart et al. (2008) consider the delta region of Bangladesh as only moderately vulnerable as a result of investments in preparedness and risk reduction, including the establishment of early warning systems, and a strengthened response capacity.

Box 3.1. Projections of sea-level rise and water supply for Asia and Pacific countries

Countries in South Asia, Southeast Asia, and the Pacific Islands are highly vulnerable to sea-level rise, which will increase the risk of floods. The global sea level gradually rose during the 20th century and continues to rise at increasing rates (Cruz et al. 2007). In the Asia and Pacific region, sea level is expected to rise approximately 3–16 centimeters (cm) by 2030 and 7–50 cm by 2070 in conjunction with regional sea-level variability (Preston et al. 2006).

Under a conservative sea-level rise scenario of 40 cm between today and the end of 21st century, the number of people facing floods in coastal areas annually will rise from 13 million to 94 million, with

11 Risk hotspots combine hazard risks with human vulnerability.

Page 58: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

58

60 percent of this increase occurring in South Asia (coasts of Bangladesh, Burma, India, Pakistan, and Sri Lanka) and 20 percent in Southeast Asia (coasts of Indonesia, the Philippines, Thailand, and Viet Nam) (Cruz et al. 2007).

Studies on the vulnerability of coastal zones to sea-level rise and storm surges are severely hampered by lack of data on coastal protection, including both natural and man-made protection systems. It is likely, however, that the low-lying river deltas of Bangladesh, China, India, Viet Nam, and the small island states in the Pacific face the biggest risk of coastal inundation, soil erosion, displacement of communities, loss of agricultural land, intrusion of saline waters into surface and groundwater, and other consequences of sea-level rise (Parry et al. 2004; Arnell et al. 2002; Cline 2007; Preston et al. 2006; Parry, Rosenzweig, and Livermore 2005; Preston et al. 2006). In the Zhujiang Estuary in China, for instance, sea-level rises of 0.4 to 1.0 m can induce further saltwater intrusion of one to three km (Bates et al. 2008). Although this particular distance is quite small, such distances can be significant if they interrupt domestic or irrigation water supplies.

The most dominant climatic drivers for water availability are precipitation, temperature, and evaporative demand (Kundzewicz et al. 2007). Although climate change affects the volume and timing of river flows and groundwater recharge, greater water demand in the future as a result of population and economic growth outweighs climate change in defining the state of future global water systems (Vorosmarty et al. 2000; Arnell 2004). The impacts of climate change, however, will continue to increase in importance over time. Scenarios show that in parts of South and East Asia, climate change will increase runoff, which is likely to increase the risk of floods during the wet season, while Central Asia will face a decrease in mean runoff (Arnell 2004; Warren et al. 2006; Shrestha and Yatsuka 2008). In the Mekong, the maximum monthly flow is projected to increase by 35–41 percent in the basin and by 16–19 percent in the delta (by 2070–99 compared with 1961–90 levels), and the minimum monthly flow is projected to decline by 17–24 percent in the basin and 26–29 percent in the delta. The expected results are increased flooding risk during the wet season and water shortages in the dry season (Bates et al. 2008). In arid and semi-arid Central Asia, climate change is expected to increase the challenges countries face in meeting growing demand for water (Bates et al. 2008).

Climate change is also likely to affect groundwater resources by altering recharge capacities in some areas, increasing demand for groundwater as a result of less surface water availability, and causing water contamination due to sea-level rise (Shrestha and Yatsuka 2008). In Asia, around 2 billion people depend on groundwater resources for drinking water, and agriculture is the largest user of groundwater resources. Agricultural systems are highly dependent on groundwater resources in India (60 percent of total agricultural water use), Pakistan’s Punjab (40 percent), and the Shangdong, Henan, Beijing, and Hubei provinces of China (50 percent, 50 percent, 65 percent, and 70 percent, respectively, of total water use) (Shrestha and Yatsuka 2008).

Pakistan

Pakistan is another country expected to be extremely vulnerable to climate change under all scenarios by 2100 (see Figure 3.2) (Yohe et al. 2006). Agriculture contributes about 20 percent of total GDP and employs more than 40 percent of the total workforce (World Bank 2008), making the country sensitive to global warming. Around 23 percent of the population lives below a poverty line of US$1.25 (Bauer et al. 2008), which directly affects communities’ ability to cope with climate change. Furthermore, only one-fourth of the country’s land is arable, and 80 percent of this land depends on irrigation and faces serious land and water degradation (O'Brien 2000).

Ehrhart et al. (2008) consider most of Pakistan to face high human vulnerability with both flood and drought hot spots—an exception is the Indus basin with fertile land and ample water supply. Brooks, Adger and Kelly (2005) go even further reporting on high climate-related

Page 59: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

59

mortality in the country associated with poor outcomes for several health, governance, and education indicators.

Box 3.2 Glaciers in the Himalayas and Central Asia are already melting as a result of global warming

Himalayan glaciers form a reservoir that supports perennial rivers on which millions of people in Bangladesh, Bhutan, India, Nepal, and Pakistan depend for survival (Cruz et al. 2007). Around 10 percent of the volume of Himalayan rivers comes from melting water from the glaciers, which are essential to sustain river flows during dry seasons (Mirza 2007). As a result of global warming, the Himalayan glaciers are receding faster than any glaciers in the world. If the present rate of melting continues, there is a high chance that they will disappear by 2035 (Cruz et al. 2007). The Dokriani glacier, for instance, which feeds the Ganges River, receded 20 meters in 1998, compared with an annual average of 16.5 meters from 1993 to 1998, and the Gangotri glacier, which receded at an annual average of 7.3 meters from 1842 to 1935, receded 23 meters a year from 1985 to 2001 (Mirza 2007; Cruz et al. 2007). Mirza (2007) reports on some of the implications of the melting of the Himalayan glaciers: more water in the perennial rivers in the Himalayas in the short run, which might be positive during dry seasons but might also increase the chance of floods (from glacial lake outbursts, for example). The short-term increase in dry-season flows might also increase sediment supply in the rivers, which may pose a threat to dams and reservoirs in the region. In the long run, declines in dry-season flows to below current levels are likely, however, posing threats to food security and the environment (Mirza 2007; Preston et al. 2006). In Central Asia, a region highly dependent on irrigation, glacier melt has increased substantially since the 1970s. In Tajikistan, for instance, glaciers lost a third of their area in the second half of the 20th century. As in the Himalayas, the melting of glaciers is expected to increase flows in Central Asia in the short run but exacerbate water shortages in the long run (Schubert et al. 2008). Moreover, rapidly melting glaciers, glacial runoff, and glacial lake outburst are already causing mudflows and avalanches in Asia (Schubert et al. 2008).

Cambodia

Cambodia is considered one of the most vulnerable countries in Southeast Asia. Although the country is not highly exposed to climate hazards, adaptive capacity is very low. Part of the country lies in the Mekong Delta (Yusuf and Francisco 2009), and highland areas are threatened by landslides. About 40 percent of the population lives on less than U$$1.25 a day, and 30 percent of GDP comes from agriculture (Bauer et al. 2008; World Bank 2008). By 2100, Cambodia is expected to be extremely vulnerable to climate change in three out of four scenarios (Yohe et al. 2006); only the combined mitigation and enhanced adaptive capacity scenario shows better outcomes for the country. In that study, however, the authors did not project climate extremes.

Viet Nam and Mekong Delta

Although Viet Nam generally has a high adaptive capacity, much of the country is subject to flood and drought risks as well as cyclones (Ehrhart et al. 2008; Yusuf and Francisco 2009). The Mekong Delta is also a hotspot for sea-level rise (Yusuf and Francisco 2009). Dasgupta et al. (2007), in a study assessing the impacts of continued sea-level rise on 84 coastal developing countries, included Viet Nam among the top five most-affected countries. A one-meter rise in sea level would affect 11 percent of Viet Nam’s population (existing population), 16 percent of its land area, and 7 percent of its agricultural area. Similarly, Cruz et al. (2007) project that a one-

Page 60: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

60

meter sea-level rise would flood 5,000 km2 of the Red River Delta and 15,000–20,000 km2 of the Mekong River Delta, affecting 4 million and 3.5–5 million people, respectively.

Mekong Delta. Global warming is also expected to affect other Mekong riparians, in particular Lao PDR and Thailand. A study of the Mekong River’s tributaries in Lao PDR and Thailand shows that climate change is likely to increase water levels in most tributaries owing to higher precipitation, which will increase the risk of flooding (Snidvongs 2006).

China and India

In countries such as China and India, structural change and growth might reduce future sensitivity to climate change and increase the ability to leverage resources to reduce risk (Preston et al. 2006). Parts of these countries, however, are still seen as highly vulnerable to climate change. According to Yohe et al. (2006), both countries are expected to be extremely or significantly vulnerable to climate change by 2100, even considering mitigation and enhanced adaptive strategies. Ehrhart et al. (2008) consider areas in India to be flood, cyclone, and drought hot spots and parts of northern and western China to be flood and drought risk hotspots, and thus subject to high human vulnerability. Furthermore, an analysis that combines indicators measuring sensitivity to climate change and adaptive capacity (but not exposure) presents India and China as vulnerable countries now and in 2095 (in two out of three scenarios) (Moss, Brenkert, and Malone 2001). In China, pasture degradation is another factor that increases the vulnerability of the agricultural sector to climate change (see Box 3.3).

Box 3.3 Pasture degradation in Inner Mongolia, China and Mongolia

Grazing areas occupy about 26 percent of the ice-free terrestrial surface of the planet. The total area occupied by feedcrop production is equivalent to 33 percent of total arable land. Livestock production accounts for 70 percent of total agricultural land and 30 percent of the land surface of the planet (Steinfeld et al. 2006). Pasture degradation, caused by a mismatch between livestock density and capacity of the pasture to be grazed and trampled, is common in semi-arid and arid areas of both Africa and Asia. Degradation can cause soil erosion, degradation of vegetation, carbon release from organic matter decomposition, loss of biodiversity, and impaired water cycles (Steinfeld et al. 2006). Grassland degradation caused by overgrazing can exacerbate the vulnerability of livestock systems to climate change. Studies have shown that grassland productivity is very sensitive to precipitation changes (Christensen et al. 2004; Chullun, Tieszen, and Ojima 1999). In Mongolia, for instance, 90 percent of rangeland area, which constitutes more than 80 percent of total area, is under threat of desertification, and degraded land has increased by 8–10 percent over the past decade (Ji 2008).

In Inner Mongolia, China, rangelands (about 67 percent of total area) have been steadily deteriorating at an annual rate of approximately 2 percent a year as a result of a combination of factors such as overgrazing (high livestock density) and climatic stress, and 55–60 percent of total area experiences desertification processes. Rangeland productivity has declined in the past five decades in meadow steppe (54–70 percent), typical steppe (30–40 percent), and desert steppe grassland areas (50 percent) (Angerer et al. 2008). Simulations in the region show that a combination of increased precipitation, temperature, and CO2 fertilization would have synergistic effects on the typical steppe grassland production of the region (Christensen et al. 2004). Herbaceous aboveground net primary production (ANPPh), however, was found to be most sensitive to changes in precipitation levels. Large decreases in precipitation caused a decline in ANPPh through a decline in soil water, which in turn decreased plant growth rates. Experiments simulating a decline in livestock density showed that declines in ANPPh were reduced or even reversed.

Page 61: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

61

Central Asia

Central Asia is a region clearly in need of more climate change–related research. Socioeconomic indicators substantially vary among the countries. People living on less than US$1.25 a day account for almost 40 percent of the population of Uzbekistan, but only 0.03 percent of the population of Azerbaijan (Bauer et al. 2008). The population of the region is highly dependent on agriculture for survival, with agriculture employing more than 30 percent of the total labor force in all countries. Georgia’s agricultural sector employs 54 percent of the total workforce (World Bank 2008).

Land and water degradation already contribute to crop yield declines in the region, which might be further exacerbated under climate change (see also Box 3.4). Moreover, Central Asian countries are heavy consumers of water for irrigation. In Uzbekistan, for instance, agriculture consumes more than 90 percent of the water used in the country (FAO 2007). Countries of the region have been consuming water at an unsustainable rate for decades, and since independence in 1991, water use has intensified even more (Allouche 2004). Furthermore, more than half of all irrigated areas in the region are salinized, waterlogged, or both. About two-thirds of land area in Kazakhstan is affected by desertification. The area is even higher in Turkmenistan and Uzbekistan, at 80 percent. In Kyrgyzstan and Tajikistan, 88 percent and 97 percent of agricultural land, respectively, is affected by erosion (Ji 2008). Thus, land degradation, desertification, and droughts in Central Asia are common in all countries, directly affecting the region’s people, who live mostly in rural areas.

Box 3.4 Climate change and land degradation in Asia*

Around 54 million km2 or 40 percent of global land area is occupied by drylands, of which the largest share (34 percent) is in Asia. Around 25 percent of the land in Asia is vulnerable to land degradation (WMO 2005). Agroecological zone assessments indicate that 28 percent of the soils in Asia suffer severe fertility constraints, and 11 percent are affected by limitations resulting from salinity, sodicity, or gypsum constraints. Around 90 percent of very suitable and suitable rainfed land is currently cultivated, which leaves little room for expansion of agricultural area. The projected population increase in Asia—an additional 1.7 billion people by 2050—will reduce per capita availability of cultivated land to less than 0.1 ha per person (Fischer et al. 2001), increasing pressures on land that might contribute to land degradation.

Climate change is likely to increase the vulnerability of poor farmers who already struggle with land degradation. Poor farmers do not have the same access to alternative sources of income as rich farmers, such as borrowing and repaying in better years. They also lack the resources for sustainable land management to maintain yields. As a result, unsustainable practices lead to further degradation (FAO 1994).

A doubling of CO2 in the atmosphere might lead to a 17 percent increase in the world’s area of desert land (WMO 2005). Soil erosion can result not only from lack of rainfall, but also from too much rainfall, because surface runoff caused by extreme rainfall events carries soil particles away and transports agricultural chemicals, contaminating groundwater. Soil erosion will likely increase the number of landslides in the hilly areas of East and Southeast Asia. Wind erosion is another cause of land degradation. In China, wind erosion buries 210,000 ha of productive land annually, a situation that is likely to worsen given that the frequency of strong sandstorms in China has increased from 5–8 annually (in the 1950s and 1960s) to 14–20 (in the 1980s and 1990s) (WMO 2005).

In many countries, land-cover changes come at the cost of increased degradation of ecosystems. In most delta regions of Bangladesh, China, India, and Pakistan, increased aridity has already resulted in the drying of wetlands and ecosystem degradation (Bates et al. 2008). In countries such as Indonesia,

Page 62: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

62

which has the world’s third-largest area of tropical forest (15 percent of the world’s forest area), land expansion would come at a substantial environmental cost. In fact, 50 percent of forest area in Indonesia is already degraded, and some parts are in critical condition (Sari et al. 2007). In Asia as a whole, there are 25 million ha of land in forest ecosystems with rainfed cultivation potential for wheat, rice, or maize (6.5 percent of total forest land area with cultivation potential). The consequences of forest clearing would be serious, however, from loss of biodiversity to the disruption of carbon sinks, hydrological cycles, and fragile ecosystems (Fischer et al. 2001). Note: This box does not include the Pacific Island countries because of a lack of available data.

Pacific Island Countries

Indicators for exposure, sensitivity, and adaptive capacity to climate change for the Pacific Island region have not been studied in detail. For instance, there is a lack of reliable poverty data for the region. Work by the Asian Development Bank (ADB) and the United Nations Environment Programme (UNDP), however, shows that poverty is increasing in those countries (Yari 2003). Small islands in the Pacific are particularly vulnerable to sea-level rise because the region is the center of ENSO. Sea-level-rise data older than 50 years from four stations in the Pacific region reveal that the average rate of sea-level rise in the region is 0.16 cm a year. Twenty-two stations with data older than 25 years estimate an average rate of relative sea-level rise of 0.07 cm a year (Bindoff et al. 2007)12. A study by the World Bank suggests that under a best-guess scenario, 18 percent of Buariki, an island in Kiribati, could be inundated by 2050, and 30 percent by 2100. If storms surges are included in the scenarios, up to 80 percent of Buariki could be inundated by 2050 (World Bank 2000).

Several current vulnerabilities of the island states are likely to make the impacts of climate change, particularly extreme events and sea-level rise, more intense, threatening food security in these countries. In Vanuatu, for instance, small farms are scattered across the islands, which makes it extremely difficult to provide services to farmers. During natural disasters, access to farms becomes even more difficult, which affects agricultural production and trade and consequently the food security of the country (FAO 2008). In the Marshall Islands, a high population growth rate puts considerable pressure on water and land resources, which increases food insecurity caused by climate change (FAO 2008). In Timor Leste, increases in extreme weather events will affect a population that is highly food-insecure and dependent on subsistence agriculture (Reske-Nielsen 2008).

In many Pacific Island countries, farmers are increasingly growing nontraditional crops that can grow in poor soils and require low labor inputs (World Bank 2000). These crops, however, are less resilient to the tropical cyclones that occur regularly in the region. The combination of more intense cyclones and the trend toward cultivation of nontraditional crops will result in greater food crop losses than would occur if traditional root crops were maintained (World Bank 2000). A recent assessment carried out in four Pacific countries (Fiji, Papua New Guinea, Tonga, and Vanuatu) showed that food security systems in rural areas are mainly based on natural resources, whereas urban areas are more dependent on imported food (FAO 2008). In these countries, the poorest and most vulnerable segments of the population, such as those dependent on subsistence fisheries and crops, are likely to be the most affected by climate change (World Bank 2000).

12 The authors of the study mention that data sets contain a large range of rates of relative sea level change, presumably as a

result of poorly quantified vertical land motions.

Page 63: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

63

3.3 Vulnerability indicator for Asia and the Pacific region

A simple, but consistent vulnerability indicator can be constructed by combining elements of exposure to climate change, sensitivity to climate change, and adaptive capacity. Results are presented in Table 3.3 and Figure 3.4. Exposure was reflected as the delta change in temperature and annual precipitation in 2080 as compared with current levels (average of 1961–90, Table 1.3). Countries were classified as highly exposed if the temperature is expected to increase by at least 2 degrees or if annual precipitation levels are projected to change by at least 20 percent, using results from the Hadley A2a scenario. Data were not available for several Pacific islands. As mentioned in Chapter 2, how those changes will affect agriculture and livestock production in the Asia and Pacific countries will depend on several factors such as crop type, CO2 fertilization, and multiple stressors. The second element of vulnerability—sensitivity—may be assessed through several variables. For instance, in the Asia and Pacific region, many countries are sensitive to climate change and climate extremes because of the high dependency of their economy on agriculture, high water stress, and land degradation rates. Other countries have low-lying coastal areas that are more sensitive to the impacts of sea-level rise and storm surges (Preston et al. 2006). Therefore, many indicators can be used to assess the sensitivity of countries’ agriculture to climate change, such as rural population density, irrigated land, and agricultural employment. In this case, sensitivity was represented by the share of labor employed in agriculture (FAO 2004). Countries with agricultural employment above 40 percent were considered highly sensitive. Bhutan, Nepal, and Timor-Leste have the highest rates of agricultural employment in the region as a share of total employment, all above 80 percent. On the other hand, the Republic of Korea and Singapore have less than 10 percent of the labor force working in agriculture.

Likewise, several indicators can be used to measure adaptive capacity, such as poverty rates, access to credit, literacy rates, farm income, and agricultural GDP. In this case the level of poverty was used to represent adaptive capacity in the Asia and Pacific region (Bauer et al. 2008). A poverty level of more than 30 percent was considered to indicate low adaptive capacity, although it is important to remember that projections of future levels of adaptive capacity are highly uncertain (Patt, Klein, and Vega-Leinert 2005). Annex 3.1 presents the indicator component data for sensitivity; Annex 3.2 presents the indicator data for adaptive capacity; and Table 1.3 for exposure for the countries in the Asia and Pacific region.

A combination of these three indicator components identifies Afghanistan, Bangladesh, Cambodia, India, Lao PDR, Myanmar and Nepal as most vulnerable to climate change—with poor outcomes in all three vulnerability components—revealing South and Southeast Asia as the regions most vulnerable to climate change. Although some of the adaptation (and mitigation) responses will be similar for all four South Asian countries, significant differences will likely be needed to adapt to climate change in Afghanistan versus Bangladesh, for example. Details on adaptation options are presented in Chapter 5. Countries with significant vulnerability—poor outcomes in two out of three components—include Bhutan, China, Pakistan, Thailand, Timor-Leste, Uzbekistan and Viet Nam, which are scattered throughout the Asia and Pacific region. The vulnerability indicator does not include the role of climate extremes and also neglects sea-level rise and glacier melt—two climate change–related events affecting the Asia and Pacific region in particular. Countries highly vulnerable to natural disasters—Bangladesh, India, and Viet Nam as well as some of the island states—made the climate change vulnerability list

Page 64: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

64

without specific inclusion of sea-level rise or glacier melt. However, the general lack of data on the Pacific Islands might neglect their relative vulnerability levels.

3.4 Conclusion

Vulnerability to climate change depends not only on exposure to climate events, but also on physical, environmental, socioeconomic, and political factors that influence how sensitive countries will be to a changing climate and how they will be able to cope.

Studies show that several countries in the Asia and Pacific region have high levels of vulnerability to climate change exacerbated by low adaptive capacity. South and Southeast Asia are among the most vulnerable regions in the world to the impacts of extreme events. Countries in South Asia, Southeast Asia, and the Pacific Islands are highly vulnerable to sea-level rise, which will increase the risk of floods. Climate change will likely increase runoff in parts of South and East Asia, whereas runoff in Central Asia is expected to decline. In the Mekong Delta in particular, increased variation in flows will likely aggravate risks of floods and droughts.

Glaciers in the Himalayas and Central Asia are already melting as a result of global warming. This development has potential short-term benefits and risks but will likely have adverse long-term impacts on food production and ecosystem health in the dry season. Climate change is also likely to increase vulnerability of poor farmers who already struggle with land degradation. In regions highly dependent on livestock production, such as Mongolia and Inner Mongolia, China, overgrazing increases vulnerability to climate change.

The countries most vulnerable to climate change are Afghanistan, Bangladesh, Cambodia India, Lao PDR, Myanmar and Nepal, and countries with significant vulnerability include Bhutan, China, Pakistan, Thailand, Timor-Leste, Uzbekistan and Viet Nam. As in Africa, those countries least to blame for climate change are likely to suffer most from its adverse impacts as a result of their location and low adaptive capacities. As shown by the improved resiliency of Bangladesh to withstand tropical cyclones in 1997 as compared with 1991, however, adaptation is possible even for the most destitute and vulnerable countries.

Each of the three components defining vulnerability to climate change—exposure, sensitivity, and adaptive capacity—requires several strategies in order to reduce the vulnerability of agriculture and rural communities in the Asia and Pacific region. Mitigation and adaptation measures are essential as a way to reduce the extent of global warming, to reduce the sensitivity of countries, and to improve the capacity of countries to adapt to a changing climate.

Vulnerability assessments are important to ensure that scarce public and private resources are allocated to those most in need of adapting to climate change. Although various vulnerability assessments generally come to similar conclusions, differences in results do exist because of the use of different data, different factors representing vulnerability, and differing methodologies. Care must therefore be taken when drawing further conclusions or basing investment decisions on such assessments.

Page 65: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

65

CHAPTER 4. OPPORTUNITIES FOR MITIGATION AND SYNERGIES WITH ADAPTATION AND SUSTAINABLE DEVELOPMENT

Key Messages

Asia is a key emitter of GHG through fertilizers and soils (N2O) as well as livestock and rice production (CH4). Much of the expected emission increase in agriculture will be in Asia as a result of food production growth required to feed a larger, wealthier population.

Agriculture has a significant technical mitigation potential (5.5–6 Gt CO2-eq by 2030), which can be implemented using zero- and low-cost technologies. Up to 50 percent of the global theoretical mitigation potential could be realized in Asia. The potential is particularly high in Southeast Asia.

Asia could potentially reduce emissions by 276.79 Mt CO2-eq a year at a carbon price of US$20/t CO2-eq--approximately 18 percent of total global economic potential (including soil carbon sequestration). At this price, the benefit stream from mitigation in agriculture for Asia could amount to more than US$5.5 billion a year. However, much of the low-cost economic potential is located outside of Asia.

Key GHG low- or no-cost mitigation activities in the Asia and Pacific region include low or no till and other sequestration methods, as well as reducing methane emissions from rice fields. China and India could each reduce methane emissions from rice fields by 26 percent over the baseline at low cost (less than US$15/t CO2-eq) by 2020. Using high-yielding varieties, shifting to rice-wheat production systems, and alternating dry-wet irrigation are technologies that both mitigate emissions and build resilience by conserving water, reducing land requirements, and reducing fossil-fuel use.

There is significant potential for small farmers to sequester soil carbon if appropriate policy reforms are implemented. If the high transaction costs for small-scale projects can be eliminated, carbon markets could be a significant source of financing. Successful implementation of soil carbon trading would generate significant co-benefits for soil fertility and for long-term agricultural productivity.

As with adaptation, the outcome of international climate change negotiations will have major effects on the role of agriculture in mitigation. Actions toward including agriculture into a post-Kyoto regime have to be taken now with a focus towards integrating smallholder farmers in carbon markets.

Synergies between adaptation and mitigation strategies need to be exploited. They can help build ecosystem resilience and generate income helping to ensure food security in the subregions of Asia and the Pacific.

4.1 Introduction

Mitigation and adaptation are both essential to deal with climate change. But adaptation becomes costlier and less effective as the magnitude of climate change grows, so the more mitigation that can be achieved at an affordable cost, the smaller the burden placed on adaptation and the lesser the suffering. As discussed in Chapter 3, scenarios that include both mitigation and adaptation to the future effects of climate change result in lower levels of vulnerability than scenarios with

Page 66: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

66

mitigation or adaptation alone (for example, see Yohe et al. 2006, Figure 3.2). Therefore, pursuing synergies between mitigation and adaptation in the context of poverty reduction will particularly build resilience against the effects of climate change.

Global technical mitigation potential in the agricultural sector is high, estimated at between 5.5 and 6.0 gigatons (Gt) of CO2 equivalent (CO2-eq) per year by 2030, with a potential for Asia to contribute up to 50 percent of theoretical reductions (Smith et al. 2007a, calculated from Figure 4.1). Box 4.1 describes the terms theoretical/technical and economic mitigation potential further. The majority of these emission reductions can be achieved through effective changes in agricultural management practices that increase soil carbon, reduce methane emissions from flooded rice fields, and improve nitrogen fertilizer usage.

Nearly 60 percent of the population of the Asia and Pacific region depend on agriculture and therefore have the potential to contribute to effective emissions reduction strategies. Moreover, adopting effective management practices that reduce greenhouse gas (GHG) emission will have significant co-benefits with adaptation and provide additional livelihood strategies (FAO 2009; Bryan et al. 2008). Finally, with the establishment of functioning carbon markets, mitigation strategies in the agricultural sector have the potential to generate financial flows to the region, potentially generating income in rural areas, and thereby increasing adaptive capacity.

4.2 Global and regional emissions trends and sources

Agricultural activities release significant amounts of GHGs into the atmosphere. The share of agricultural emissions in total GHG emissions was 13 percent in 2000, and if land use change is considered, agriculture contributes upwards of 30 percent of global emissions (Figure 4.2). Emissions from this sector are primarily methane (CH4) and nitrous oxide (N2O), making the agricultural sector the largest producer of non-CO2 emissions, with 60 percent of the world total in 2000 (WRI 2008). Although agricultural lands also generate large CO2 fluxes both to and from the atmosphere through photosynthesis and respiration, this flux is nearly balanced on existing agriculture lands. Significant carbon release, however, results from the conversion of forested land, which is included in the category of land use, land use change, and forestry (LULUCF).13 Finally, other agricultural activities related to GHG emissions are included in other sectors, such as the upstream manufacture of equipment, fertilizers, and pesticides; the on-farm use of fuels; and the transport of agricultural products.

Regional variations in emissions from agricultural sources (non-CO2) indicate that countries outside the Organization for Economic Cooperation and Development (OECD) emit nearly 75 percent of global emissions (WRI 2008). As a result, the theoretical potential for mitigation in the agricultural sector is greater in developing countries than in industrial countries. Asian countries account for 37 percent of total world emissions from agricultural production, China alone accounts for more than 18 percent of the total (WRI 2008).

Emissions from agriculture come from four principal sectors: agricultural soils, livestock and manure management, rice cultivation, and the burning of agricultural residues and savanna

13 Total LULUCF emissions, which include biomass clearing and burning for agriculture and urban expansion, as well as timber

and fuelwood harvesting, were nearly 18 percent of total GHG emissions in 2000, or 7,618 Mt CO2 equivalents. Concerning food production specifically, estimates of the amount of total emissions in this sector that are due to land conversion for agricultural extensification are difficult to make. One estimate, however, attributes 9 percent of total global emissions—one-half of LULUCF emissions—to expansion into forests for feedcrop and livestock production (Steinfeld et al. 2006).

Page 67: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

67

for land clearing. Figure 4.3 presents the share and pollutants from each of these sectors. The largest shares of emissions originate from agricultural soils (N2O), enteric fermentation associated with livestock production, and rice production (CH4). Emissions from agriculture are expected to rise because of increased demand for agricultural production for growing and more prosperous populations able to afford more varied diets with higher shares of meat and dairy products (see, for example, Delgado et al. 1999). This shift will also lead to increased pressure on forests from agricultural expansion. As a result, both emissions from fertilizers and livestock are expected to continue to increase significantly out to 2020 (Figure 4.4).

South and East Asia contribute 43 percent of global N2O emissions from soils (1,136 Mt [megatons] of CO2 equivalent a year, Table 4.1). East Asia alone emits 68 percent of total global CH4 from rice production, and South Asia accounts for another 20 percent of the global total of 561 Mt CO2 equivalent a year. The release of CH4 from enteric fermentation from these two subregions emits an additional 569 Mt CO2 equivalent a year, or 47 percent of the global total in this category. Fertilizer and manure applied on soils were the main sources of N2O, whereas the large livestock population contributed to the high enteric fermentation that releases CH4 gases (USEPA 2006). Emission trends in these two subregions will continue to 2020, when N2O from soils is expected to nearly double to approximately 2,000 Mt CO2 equivalent a year, methane emissions are expected to increase to 1,250 Mt CO2 equivalent a year, and methane emissions from livestock will rise by a third to approximately 800 Mt CO2 equivalent a year (Figure 4.5). Overall in 2020, the developing countries of East and South Asia are expected to emit 2,800 Mt CO2 equivalent a year across all agricultural sources.

4.2.1 Agricultural soils

Nitrous oxide is the largest source of GHG emissions from agriculture, accounting for 38 percent of agricultural GHGs globally. N2O is produced naturally in soils through the processes of nitrification and denitrification. Activities may add nitrogen to soils either directly or indirectly. Direct additions occur through nitrogen fertilizer use, application of managed livestock manure and sewage sludge, production of nitrogen-fixing crops and forages, retention of crop residues, and cultivation soils with high organic-matter content. Indirect emissions occur through volatilization and subsequent atmospheric deposition of applied nitrogen, as well as through surface runoff and leaching of applied nitrogen into groundwater and surface water (USEPA 2006).

Direct application of nitrogen-based fertilizers, both synthetic and organic, is a major source of growth in N2O emissions. Under a business-as-usual scenario, these emissions are expected to increase by 47 percent from 1990 to 2020. In 1990, the OECD and China accounted for approximately 50 percent of all N2O emissions from agricultural soils. Projections to 2020, however, show that emissions will remain relatively static in the OECD, with major increases coming from China (50 percent increase). On the other hand, current agricultural production in Central Asia is only about 60–80 percent of its 1990 level but is expected to increase by 15–20 percent above 2001 levels by 2010 owing to the increasing wealth of the countries in the region (Smith et al. 2007a). Central Asia focuses on agricultural expansion that includes an increase of 10–14 percent in arable land for the whole of Russia, as well as use of intensive management technologies in farm areas. Such trends suggest that an extensive application of production technologies will be required to achieve the 2- to 2.5-fold increase in grain and fodder yields with a consequent reduction of arable land and intensified nitrogen fertilizer usage (Smith et al.

Page 68: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

68

2007a). Furthermore, during the 1990s, lower rates of application of nitrogen fertilizer led to significant reduction of N2O emissions (Smith et al. 2007a). Given the current economic situation in Central Asia, however—as discussed in Chapter 2—the need to apply nitrogen fertilizer is inevitable. The sharpest increase in fertilizer application is expected in developing countries, which are expected to use 36 million tons more than developed countries by 2020 (Bumb and Baanante 1996).

4.2.2 Livestock and manure management

Enteric fermentation—or the natural digestive processes in ruminants, such as cattle and sheep—is the second-largest source of total emissions from agriculture, at 34 percent of the total. Other domesticated animals, such as swine, poultry, and horses, also emit CH4 as a byproduct of enteric fermentation. Manure management includes the handling, storage, and treatment of manure and accounts for 7 percent of agricultural emissions. Methane is produced from the anaerobic breakdown of manure, while nitrous oxide results from handling the manure aerobically (nitrification) and then anaerobically (denitrification), and is often enhanced when available nitrogen exceeds plant requirements (Oenema et al. 2005; Smith and Conen 2004).

Demand for beef and dairy products is expected to rise globally, with sharp increases in consumption and production in the developing world. By 2020, more than 60 percent of meat and milk consumption will take place in the developing world, and the total global production of beef, meat, poultry, pork, and milk will at least double from 1993 levels (Delgado et al. 1999). As a result, CH4 emissions from enteric fermentation are projected to increase 32 percent by 2020, with China, India, and Pakistan as the top sources (Figure 4.5). In addition, CH4 and N2O emissions from manure management are expected to increase an estimated 21 percent, with large shares from China.

4.2.3 Rice cultivation

Flooded rice fields are the third-largest source of agricultural emissions, contributing 11 percent in the form of CH4, which results from anaerobic decomposition of organic matter in an oxygen-deficit environment (Mosier et al. 1998). China and Southeast Asian countries produce the lion’s share of methane emissions from rice, accounting for more than 90 percent of the total in 1990. Because of population growth in these countries, emissions are expected to increase 36 percent in Southeast Asia and 10 percent in China by 2020 (USEPA 2006).

4.3 Mitigation strategies in the agricultural sector

As outlined here, the biological processes associated with agriculture are natural sources of GHGs. Yet farmers have the potential to reduce the quantity of emissions through the efficient management of carbon and nitrogen flows. Mitigation is a response strategy to global climate change, defined as measures that reduce the amount of emissions (abatement) or enhance the absorption of GHGs (sequestration). The total global potential for mitigation depends on many factors, including emissions levels, technology availability, enforcement, and incentives. In many situations, agricultural efficiency can be improved at a low cost; however, when low-cost incentives are unavailable, policy development is important.

Mitigation strategies in agriculture can be categorized in three ways: carbon sequestration into soils, on-farm emission reductions, and emission displacements from the transportation

Page 69: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

69

sectors through biofuel production (Smith et al. 2007a). These three options for mitigation in agriculture will be further discussed.

4.3.1. Carbon sequestration

Sequestration activities enhance and preserve carbon sinks and include any practices that store carbon through cropland management “best practices,” such as no-till agriculture, or that slow the amount of stored carbon released into the atmosphere through burning, tillage, and soil erosion. Sequestered carbon is stored in soils, resulting in increases in soil organic carbon (SOC). SOC will approach a new equilibrium over a 30–50 year period, however, and is therefore limited by saturation. In addition, there is potential for the re-release of SOC into the atmosphere through fire or tillage, which raises concerns about the “permanence” of SOC storage. On the other hand, emissions abatement through improved farm management practices could be sustained indefinitely. Despite these limitations, soil carbon sequestration is estimated to account for 89 percent of the technical mitigation potential in agriculture, compared with 11 percent for emissions abatement (Smith et al. 2007a). Figure 4.6 shows the dominance of soil carbon sequestration (CO2) in technical mitigation potential.

Many best management practices in agriculture raise SOC. Such practices include reducing the amount of bare fallow, restoring degraded soils, improving pastures and grazing land, adopting irrigation, rotating crop and forage, and practicing no-till agriculture (Smith et al. 2007a). The total technical potential of global cropland soils to sequester carbon through a combination of these techniques has been estimated at 0.75 to 1 Gt a year (Lal and Bruce 1999). Specifically the South Asia region could increase SOC by 25 to 50 teragrams (Tg) C a year (1 Tg = 1 Mt) through conversion to restorative land use and adoption of integrated nutrient management14 (Lal 2004). One technique that the literature highlights as having a high mitigation potential is no-till agriculture (see Box 5.1). Estimates show that tillage reductions on

14 Lal (2004) considers integrated nutrient management practices such as “use of manure, compost, green manuring and other

biosolids including city sludge, mulch farming, conservation tillage, and diverse crop rotations based on legumes and cover crops in the rotation cycle.” In addition, the use of chemical fertilizer also increased the formation of SOC in combination with integrated nutrient management practices (Lal 2004).

Box 4.1. Technical versus economic mitigation potential

The technical mitigation potential is the theoretical amount of emissions that can be reduced and the amounts of carbon that can be sequestered given the full application of current technologies, without considering the costs of implementation. It describes the order of magnitude that current methods of mitigation may allow, instead of providing realistic estimates of the amount of carbon that will be reduced under current policy and economic conditions.

Calculations of economic mitigation potential consider the costs of technologies as well as the cost of carbon over a price range. The economic mitigation potential of technologies can be analyzed by determining the marginal abatement cost (MAC) or the cost of reducing emissions by one unit. These costs can be plotted over varying price levels in order to show the relationship between carbon price and the percentage reduction in emissions.

Both the estimates of technical and economic potential need to be treated with some caution. In general, they do not consider trade-offs with other goals, such as income generation or food security, nor do they consider the heterogeneity in management capacity or cultural appropriateness.

Page 70: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

70

global cropland could provide a full “wedge” of emissions reductions—up to 25 Gt over the next 50 years (Pacala and Socolow 2004). Some researchers, however, have noted that tillage reductions may not be feasible in all soil types (Chan, Heenan, and So 2003). Baker et al. (2007), however, argue that improper sampling techniques together with modern gas-based measurements cast doubt on previous findings of positive carbon offsets through tillage reductions.

Another way to increase SOC is grazing land management, which increases the cover of high-productivity grasses and overall grazing intensity. Degraded or overgrazed land can be restored to produce more biomass by selectively planting grasses, adding phosphate fertilizers, and alternating grazing with rest periods for the land (Robert 2001). Increasing biomass productivity on grazing lands enhances soil cover, increases moisture availability, and increases the overall amounts of stable organic matter in the soil. In Asia, large technical potential exists in India, which has one of the world’s largest grazing land areas.

4.3.2 Bioenergy

The production of liquid fuels from dedicated energy crops, such as grains and oilseeds, is being evaluated for use as transportation fuel in response to concerns over the environmental sustainability of continued fossil fuel dependence. The potential of biofuels to reduce carbon emissions, however, is highly dependent on the nature of the production process through which they are manufactured and cultivated. There tends to be a high degree of variance in the literature over the net carbon balance of various biofuels because of differences in the technological assumptions used when evaluating the processes embedded in any life-cycle assessment. Early life-cycle assessments of biofuels found a net carbon benefit, which has contributed to consumer acceptance (for example, Wang, Saricks, and Santini 1999). Yet a number of studies are challenging the net carbon benefit in comparison with traditional fossil fuels (Pimentel and Patzek 2005), especially when biofuel production requires land conversion from cover with a high carbon sequestration value, such as forests (Searchinger et al. 2008).

Considering the impact that continued crop cover would have for agricultural soil emissions, bioenergy production is estimated to have a technical potential of approximately 200 Mt CO2-eq a year in 2030 (Figure 4.6). But the potential for GHG savings is much higher when the offsetting potential from displacement of fossil fuels is considered. It is estimated that 5–30 percent of cumulative carbon emissions would be abated if bioenergy supplied 10–25 percent of world global energy in 2030 (Ferrentino 2007). But rapid expansion in bioenergy of this magnitude would have significant trade-offs with food security, as has already been seen in the past few years, and would have significant negative impacts on food production and biodiversity (see Box 4.2). A careful assessment of these trade-offs, as well as of net GHG gains including land use change effects, needs to be undertaken for alternative bioenergy technologies as they develop.

Box 4.2. Opportunities for pro-POOR biofuel production

Opportunities for increasing the welfare benefits to the poor may arise through the use of small-scale biofuel production models that convert feedstocks locally (Ewing and Msangi 2008). Examples of small-scale production models found in the literature demonstrate a wide range of welfare gains, including new sources of energy and electricity and the development of enterprises related to coproducts, such as soap and organic fertilizer. Energy crops can be converted into fuels to satisfy a number of rural applications,

Page 71: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

71

including electrification, small machinery power, irrigation pumping, and food processing. In addition, bioenergy development for clean-burning cooking fuel, such as ethanol-based gelfuels, can provide significant time savings for women and children by reducing the need to search for and collect fuelwood. In addition, the use of clean-burning ethanol has positive health impacts, reducing the level of indoor air pollution and related illness.

Small-scale production models can also minimize food security impacts by focusing on non-edible energy crops that can be grown on marginal lands. Biofuel production on marginal land may be particularly well suited to poor farmers who do not have access to high-quality lands (Binns 2007). One crop well suited to areas with low rainfall and low soil quality is jatropha. This crop is currently being piloted in a number of small-scale biodiesel development projects in Sub-Saharan Africa and India and is the focus of a number of case studies reviewed later in this chapter. Sweet sorghum is also ideal for drier areas and has properties similar to sugarcane in ethanol production. In addition, declining demand for sweet sorghum as food, as well as its coproduction value as a livestock feedcake, lessen its threat to food security (ICRISAT 2007). A final promising variety similar to jatropha is pongamia. Although there are fewer case studies on pongamia production, this tree has been found to produce more than twice as much oil per hectare in comparison with jatropha (Rajagopal and Zilberman 2007).

Despite these benefits, there are barriers to small-scale bioenergy development in rural areas. A considerable level of effort would need to be put into the conscious design of production systems such that smallholders can directly benefit from the opportunities that biofuels may offer to the agricultural sector. At the local level, lack of technical know-how related to feedstocks and conversion, low capital availability for start-up costs, lack of private sector capacity and support, poor market development, and insecure land tenure are often cited as limitations to small-scale agricultural development. In addition, a common critique of jatropha-focused biofuel production is its rather low yield if it is grown on marginal lands without irrigation, which makes it less cost-competitive than fossil-based fuels. Most industrial processes require economies of scale and high levels of extraction efficiency to remain economically competitive, which raises the question of whether small-scale jatropha can survive in the long term without subsidies in the form of producer credits or protective tariffs on competing products.

Despite these challenges, a number of small-scale biofuel production projects across Africa and Asia are providing examples and generating knowledge of the possibilities and constraints surrounding sector development. In India, a large-scale public-private partnership has been launched to promote the profitable participation of small-scale famers in the cultivation of sweet sorghum feedstocks for ethanol production. A private business partner—Rusni Distilleries—is providing farmers with sweet sorghum seeds and feedstock supply contracts to local processing facilities in order to create a village-based supply chain model (Binns 2007). The operation of the refinery for sweet sorghum is creating 40,000 person-days of labor (ICRISAT 2007). Also in rural India, a women-led pongamia oil project used to run small generators for household electricity is being replicated by the state government in nearly 100 villages (ICRISAT 2007).

4.3.3. On-farm mitigation

Improved management practices that reduce on-farm emissions include livestock and manure management, fertilizer management, and improved rice cultivation.

Enteric fermentation

Methods to reduce methane emissions from enteric fermentation include improving digestive efficiency with improved feeding practices and dietary additives. The efficacy of these methods depends on the quality of the feed, livestock breed and age, and whether the livestock is grazing or stall fed. Developing countries are assumed to provide lower-quality feed to livestock, which raises the emissions rate per animal over developed-country herds. The technical potential to mitigate livestock emissions in 2030 is 300 Mt CO2-eq a year (Figure 4.6). Furthermore, Smith et

Page 72: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

72

al. (2007a) quantify the technical potential for reducing CH4 production through improved feed practices, specific agents and dietary additives, and longer-term structural and management changes and animal breeding (Table 4.2).

Manure management

In manure management, cooling and using solid covers for storage tanks and lagoons, separating solids from slurry, and capturing the methane emitted are effective techniques. In developing countries, however, applying this type of manure management may be difficult, as animal excretion happens in the field. Composting manure and altering feeding practices may help reduce emissions to a certain extent. The technical potential of improved manure management in 2030 is 75 Mt CO2-eq a year (Figure 4.6).

Fertilizer management

Improving the efficiency of fertilizer application or switching to organic production can decrease the nutrient load and N2O emissions. Overall benefits will need to be weighed, however, against potential impacts on yield. Although some studies (such as Delate et al. 2003; Pimentel et al. 2005) have shown that organic agriculture offers yields competitive with conventional fertilizer applications, fertilizer reductions of 90 percent in rainfed maize fields were shown to reduce yields by 10.5 percent over the baseline in China (USEPA 2006). In addition, the lack of access to soil nutrients to improve the quality of degraded soils in many parts of the developing world is a hindrance to achieving food security (Gruhn, Goletti, and Yudelman 2000). Overall, cropland management could reduce emissions in 2030 up to 150 Mt CO2-eq a year (Figure 4.6).

Rice cultivation

Improving water management in high-emitting, irrigated rice systems through midseason drainage or alternate wetting and drying has shown substantial reductions in CH4 emissions in Asia. These effects may be partially offset, however, by an increase in the amount of N2O emitted (Wassman, Butterbach-Bahl, and Doberman 2006). The technical potential of improved rice management is 300 Mt CO2-eq a year (Figure 4.6).

4.3.4 Summary of technical mitigation potential

Considering all mitigation strategies in the agricultural sector combined, the global technical mitigation potential is 5,500–6,000 Mt CO2-eq a year by 2030 (Smith et al. 2007a) (Figure 4.6). Of this estimate, carbon sequestration accounts for nearly 90 of the potential, and methane mitigation and soil N2O emission reductions account for 9 and 2 percent, respectively. Across the subregions of Asia, approximately 1,100–3,000 Mt CO2-eq a year can be mitigated by 2030 for all GHGs (Smith et al. 2007a, estimated from Figure 4.1). At the upper end, Asia could contribute 50 percent of the total technical mitigation potential by 2030.

4.3.5 Economic potential of mitigation options

Calculations of economic potential come from two main sources: Smith et al. (2007b) and USEPA (2006). The results from USEPA (2006) are preferred for non-CO2 emissions abatement because they have a finer level of regional disaggregation, which enables explicit examination of the economic potential of developing countries. Smith et al. (2007a) conducted a comparison of

Page 73: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

73

Smith et al. (2007b) and USEPA (2006) and finds consistent results across emission sources. Smith et al. (2007a, b), however, provide a more comprehensive assessment of the potential for soil carbon sequestration.

The USEPA (2006) estimates three categories of emissions mitigation and sequestration: rice cultivation; livestock and manure management; and cropland management (including N2O from fertilizer reductions, soil carbon sequestration through no tillage—but not through other management and policy changes—and split fertilization, each under both rainfed and irrigated conditions for rice, soybeans, and wheat). Marginal abatement curves are constructed for the years 2010, 2020, and 2030 to determine the relationship between carbon price and quantitative emission reductions.

Smith et al. (2007a) estimated global economic potential for agricultural mitigation using top-down and bottom-up modeling. Bottom-up mitigation responses described typical constraints to input management (such as fertilizer quantity or type of livestock feed) as well as cost estimates (partial equilibrium, where input and output market prices are fixed like acreage or production). On the other hand, the top-down mitigation responses add more generic input management responses as well as changes in output (such as shifts from cropland to forest) and market prices (such as decreases in land prices with rising production costs due to a carbon tax). Figure 4.7 presents the global estimates of economic potential for agricultural mitigation from various studies at different assumed carbon prices in 2030. Bioenergy

Neither Smith et al. (2007a) nor the USEPA (2006) calculate the marginal abatement costs of bioenergy cultivation related to agricultural soils. Estimates do, however, exist for their potential displacement of fossil fuels. Specifically for the transportation sector, liquid biofuels are predicted to reach 3 percent of demand under the baseline scenario, increasing up to 13–25 percent of demand under alternative scenarios in 2030 (IEA 2006). This outcome could reduce emissions by 1.8–2.3 Gt CO2, corresponding to between 5.6 and 6.4 percent of total emissions reductions across all sectors at carbon prices greater than US$25 per ton of CO2 (that is, US$25/tCO2) (Ferrentino 2007).

On-farm mitigation

Cropland management (N2O and CO2).Compared with the baseline, approximately 15 percent of global cropland emissions can be abated at no cost, and approximately 22 percent of emissions can mitigated for less than US$30/tCO2-eq. Beyond this point, abatement costs rise exponentially. These results are similar for all years considered. Regional calculations show that the largest zero- and low-cost potential (up to US$30/tCO2-eq) is in the Russian Federation (31.7 percent reductions over the baseline in 2020), and there is modest potential in South and South East Asia (11 percent over the baseline). The reasons that fertilizer reductions do not have a strong mitigation potential for developing countries may include existing low levels of fertilizer usage or the effect of suboptimal nutrient application on yields in some developing countries, particularly on the African continent. On the other hand, across China and India, converting from conventional tillage to no till resulted in yield increases for each crop considered. This practice thus has large potential as a negative cost option or “no-regret” scenario. Yet farmers in these regions are not adopting no-tillage practices, showing that the analysis fails to capture cost

Page 74: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

74

barriers, which may include profit variability or complex management requirements (USEPA 2006).

Smith et al. (2007a) consider a broader range of cropland management practices for soil carbon sequestration, such as reducing bare fallow and residue management. Under this broader spectrum, the economic potential for soil carbon sequestration increases up to 800 Mt CO2-eq in 2030 at carbon prices up to US$20/tCO2-eq (Figure 4.8). Given that 70 percent of total emissions abatement could come from developing countries, soil carbon sequestration will be an important management practice. Yet the economic potential of soil carbon sequestration practices in the Asia and Pacific region has not been estimated on a wide scale.

Rice cultivation. Only 3 percent of emissions from rice cultivation can be abated in 2000 at zero cost, jumping to 11 percent in 2010. Also in 2010, 22 percent of global emissions could be abated at US$30/tCO2-eq. China and India each could reduce methane emissions from rice fields by 26 percent at low cost (less than US$15/tCO2-eq) by 2020. This result is not surprising, given that China and South and Southeast Asian countries produced more than 90 percent of methane emissions from rice in 1990.

Enteric fermentation and manure management. Improved livestock and manure management together could reduce emissions by 3 percent at no cost, and between 6 and 9 percent at carbon prices of US$30/tCO2-eq. Annex 1 and OECD countries have the highest least-cost economic potential. Yet the countries with the high herd numbers, such as India, have relatively low economic potential, reducing emissions through livestock management only up to 2.5 percent at carbon prices up to US$30/tCO2-eq.

4.3.6 Summary of economic mitigation potential

Overall, opportunities for emissions mitigation in the agricultural sector at no or low cost are modest. USEPA (2006) estimates without carbon sequestration show that 9.3, 12.1, and 14.6 percent of emissions could be reduced from the baseline at carbon prices up to US$30/tCO2-eq by 2030 in India, China, and South and Southeast Asia, respectively. China and India each could reduce methane emissions from rice fields by 26 percent at low cost (less than US$15/tCO2-eq) by 2020. The consideration of expanded practices of soil carbon sequestration by Smith et al. (2007b) indicates no-tillage and other sequestration methods could have significant economic potential in Asia. Across all practices, subregions of Asia could potentially reduce emissions by 276.79 Mt CO2-eq a year at a carbon price of US$20/tCO2-eq (Smith 2009, Table 4.3). Therefore, expanding mitigation options to include potential from soil carbon sequestration expands the economic mitigation potential in Asia. At this price, investments in mitigation would amount to more than US$5.5 billion per year. Compared with total global economic mitigation potential estimated by Smith et al. (2007b), Asia could mitigate approximately 18 percent of emissions at carbon prices of US$20/tCO2-eq (calculated from Smith et al. 2007a, Figure 4.8; and Smith 2009).

4.4 Institutional barriers to mitigation in agriculture in the Asia and Pacific region

To date, little progress has been made in implementing mitigation measures in the agricultural sector at the global scale and particularly at the regional level. GHG mitigation potential would be enhanced with an appropriate international climate policy framework providing policy and economic incentives.

Page 75: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

75

The market for trading carbon emissions offers limited possibilities for agriculture to benefit from land uses that sequester carbon or save non-CO2 emissions. The Clean Development Mechanism (CDM) under the Kyoto Protocol of the United Nations Framework Convention on Climate Change (UNFCCC) is the most established mechanism for payments to developing countries. The CDM allows polluters in developed countries to purchase carbon-offsetting projects in developing countries once it has been determined that the project would not have been undertaken otherwise. Currently, eligible activities under the CDM are limited to afforestation and reforestation and reduction of non-CO2 gases in agriculture. Carbon sequestration activities, such as conservation tillage and the restoration of degraded soils, are not eligible under the CDM.

Soil carbon sequestration has the highest technical potential for mitigation in the agricultural sector, so there is room to expand agricultural sector mitigation through the CDM if carbon sequestration projects are included. Yet there are feasibility issues in selling agricultural soil carbon within a market-based credit-trading program. The transaction costs in soil carbon sequestration include the cost of obtaining site-specific information on the baseline stock of carbon and the potential to sequester carbon. The transaction costs per ton of carbon associated with negotiating contracts will decline as the size of the contract increases, and a market for carbon credits is likely to operate for large, standardized contracts (such as for 100,000 tons). For a typical individual farmer who can sequester 0.5 ton per hectare per year, these transaction costs would be prohibitive.

In addition to global mechanisms, regional institutions and financing arrangements need to be scaled up and expanded to address key region-specific climate change needs (Sharan 2008). Nascent financing arrangements have emerged to service the developing member countries of the ADB. These include the Climate Change Fund (CCF), the Clean Energy Financing Partnership Facility (CEFPF), the Asia Pacific Carbon Fund (APCF), the Future Carbon Fund (FCF), the Water Financing Partnership Facility (WFPF), and the Poverty and Environment Fund (PEF). A large share of these funds are directed toward energy efficiency, with fewer projects in biofuel, biomass and biogas promotion, and sustainable land use and forestry. Current efforts need to be expanded and scaled up to reach more farmers and broadened to include emissions reductions from soil carbon sequestration and other GHG sources in the agricultural sector.

4.5 Integrating mitigation and adaptation in sustainable development pathways

At the 2009 Delhi Sustainable Development Summit, Yvo de Boer, the executive secretary of the UNFCCC, urged that climate change policy use the opportunity to the fullest to ensure that “nationally appropriate mitigation action serves broader development goals on the one hand; and that development goals serve climate change abatement on the other.”

One of the main objectives for climate change mitigation agreements at the 15th Conference of Parties in December 2009 will be to develop “nationally appropriate mitigation actions by developing country Parties in the context of sustainable development, supported by technology and enabled by financing and capacity-building” (http://www.roadtocopenhagen.org/index.php?c=pages&id=19, accessed April 4, 2009). These initiatives reveal a growing recognition that climate change mitigation and adaptation need to be synonymous with poverty alleviation and sustainable development. Moreover, the developing member countries of the ADB have more than 2 billion people who depend on agriculture for

Page 76: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

76

their livelihoods. Therefore, mitigation and adaptation policies need to focus on agriculture and poverty alleviation together.

Current carbon financing aims to offset emissions and not to ensure pro-poor development. As a result, the bulk of projects in the CDM are aimed at large-scale emitters. Yet expanding frameworks to include the emissions reduction potential from agriculture in a post-2012 framework will help to ensure that the overall emissions reduction objective of the UNFCCC will be achieved. Moreover, developing countries, and specifically those with economies rooted mainly in agriculture or forestry, have stated that their support for a post-2012 agreement will be conditional on the inclusion of mitigation options from agriculture, forestry, and other land uses.

How can mitigation and adaptation measures be mainstreamed into sustainable development pathways? First, synergies between adaptation and mitigation should be recognized and exploited so that strategies can be mainstreamed. Second, potential economic losses that might result from pursuing synergies in adaptation and mitigation should be overcome by creating financial markets and other payments for environmental services. Lastly, it is important to ensure that carbon markets and other global, regional, and national frameworks provide adequate income flows and encourage the participation of small farmers. Each of these steps will be discussed in turn.

4.5.1 Synergies between mitigation and adaption

The strategies for reducing emissions profiled in section 4.3 also have significant synergies with adaptation. Strategies to conserve soil and water resources, such as restoring degraded soils, agroforestry, and efficient water use in rice cultivation, also enhance ecosystem functioning, increase water availability, and provide resilience against droughts, pests, and other climatic threats. In general, the mismanagement of agroecological systems generates emissions, degrades ecosystem functioning, and will ultimately threaten food security. Therefore, measures to reduce emissions through integrated crop, grazing land, pest, and water management will build ecosystem resilience, lessening sensitivity to climate change.

Rao (1994) reports that rice, nutrient, water, and tillage management can mitigate GHG emissions from agriculture. Efficient drainage and effective institutional support lower irrigation costs to farmers and thus build up the economic aspect of sustainable development. In addition, appropriate combination of rice cultivation with livestock, in what is known as an integrated annual crop-animal system, is traditionally found in India, Indonesia, and Viet Nam. This system enhances net income, improves cultivated agroecosystems, and enhances well-being (MA 2005).

Rice is the staple food and widely grown in Asia but is a significant contributor of CH4 emissions. A study by Wassmann et al. (2006) offers four approaches in offsetting CH4

emissions: (1) improving rice plants through breeding; (2) improving fertilizer management; (3) improving water management; and (4) utilizing crop residues for renewable energy and carbon sequestration. Using high-yielding varieties, shifting to rice-wheat production systems, and alternating dry-wet irrigation are technologies that both mitigate emissions and build resilience by conserving water, reducing land requirements, and reducing fossil-fuel use. In many of ADB’s developing member countries, the dependency on rice for food calories is very high (defined as greater than 800 kilocalories/person/day) (Nguyen 2005). In addition, considering that majority of the world’s rice is produced in Southeast Asian countries and India, many

Page 77: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

77

households derive their livelihoods from its cultivation. Given the strong mitigation potential estimated by USEPA (2006), where India and China could each reduce methane emissions from rice fields by 26 percent at low cost (less than US$15/tCO2-eq) by 2020, there exists potential to integrate the objectives of mitigation, adaptation, and poverty alleviation in rice production.

Improving pasture management by controlling overgrazing has favorable impacts on livestock productivity (higher income with the same number of livestock) and decelerates, if not completely minimizes, desertification (environmental aspect) (Smith et al. 2007a). In China, overgrazing is controlled by disallowing free grazing (Rao 1994). Controlling overgrazing will be a challenge, however, especially in other Central Asian countries that have large dryland and desert areas and rely on pasture grazing for food and economic needs (Smith et al. 2007a).

The breeding of improved crop varieties is another approach that has synergies with adaptation and mitigation. Crops can be bred to be more drought, pest, heat and weed tolerant and to require fewer nutrient inputs. More efficient nitrogen use by crops has several important environmental advantages, in addition to lowering production costs for farmers, in light of high fertilizer prices. Genes have been identified that improve the efficiency with which plants use nitrogen fertilizer, and genetically modified plants with these genes are currently being characterized under field conditions. The reduced need for synthetic nitrogen fertilizers will reduce energy costs and help lower GHG production.

There are barriers to creating pro-poor mitigation strategies. Often cited is the lack of technical capacity in alternative production technologies, including improved seed varieties, efficient water use, and access to extension services. In Indian rice production, methane emissions could be reduced, but only with a loss of income to farmers (Pathak and Wassmann 2007). In rapidly urbanizing regions of South Asia, income generation through brick making using topsoil compromised soil carbon sequestration activities (Lal 2004). These economic barriers need to be overcome in order to fully integrate adaptation and mitigation into sustainable development pathways. Suggestions for creating financial arrangements that address the needs of smallholders and mainstream adaptation and mitigation into livelihood strategies are given later in this chapter.

4.5.2 Suggestions for mainstreaming mitigation and adaptation in sustainable development pathways

The large scope for synergies between adaptation and mitigation practices demand that strategies be mainstreamed to maximize co-benefits. These emissions sources in the Asia and Pacific region represent opportunities for financial flows from carbon markets, payments for environmental services, and synergies with adaptation that build adaptive capacity and ecosystem integrity.

The Food and Agriculture Organization of the United Nations recommends certain features for bridging mitigation activities with broader development objectives, including:

“Financing arrangements that address specific needs in smallholder agriculture mitigation adoption including the need for investment capital and insurance”; and

“A range of options for mobilizing private and public funds for financing, including use of compliance market credits, voluntary market credits, publically funded programs and agricultural product labels (FAO 2009).”

Page 78: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

78

These recommendations underscore the need to link farmers and in particular small farmers to the financing mechanisms, technology, property rights, and capacity strengthening necessary to realize synergies between adaptation and mitigation in the context of sustainable development. The development of nationally appropriate mitigation actions (NAMAs) by developing countries is one suggestion for providing a framework for achieving these linkages which emerged under the Bali Action Plan. While not currently obliged under the Convention, these mechanisms would outline objectives and actions for technology transfers and development, capacity building, and financing needs, taking into account local development and mitigation objectives, and regional and global actions in agriculture. These plans would provide a basis for conceptualizing linkages between mitigation and sustainable development, and should be considered in the context of adaptation strategies.

In general, there is tremendous potential to link famers in developing regions of Asia and the Pacific while generating co-benefits in adaptation and sustainable development. Therefore, it is important to ensure that emerging carbon markets benefit developing countries. CDM rules should encourage the participation of small farmers and protect them against major livelihood risks, while still meeting investor needs and rigorously ensured carbon goals. CDM can support these goals by:

Promoting measures to reduce transaction costs. Rigorous but simplified procedures should be adapted to developing-country carbon offset projects. Small-scale soil carbon sequestration projects should be eligible for simplified modalities to reduce the costs of these projects. The permanence requirement for carbon sequestration should be revised to allow shorter-term contracts or contracts that pay based on the amount of carbon saved per year.

Establishing international capacity-building and advisory services. The successful promotion of soil sequestration for carbon mitigation will require investment in capacity-building and advisory services for potential investors, project designers and managers, national policymakers, and leaders of local organizations and federations (CIFOR 2002).

Investing in advanced measurement and monitoring. Proper measuring can dramatically reduce transaction costs. Measurement and monitoring techniques have been improving rapidly thanks to a growing body of field measurements and the use of statistics and computer modeling, remote sensing, global positioning systems, and geographic information systems, so that changes in stocks of carbon can now be estimated more accurately at lower cost.

Box 4.3 An example of integrating mitigation and adaptation to improve livelihoods: Biogas in Asia

Biogas produced through the anaerobic digestion of animal dung has been implemented at the household and village scale for the generation of cooking fuel, electricity, and power generation. The oldest initiative is in China, where 15 million households have access to biogas, with plans to expand biogas plants to 27 million households, or 10 percent of rural households, by 2010. Government subsidies cover up to two-thirds of the cost, and local government commitments fund the rest. Similar programs have had success in India, Nepal, and Viet Nam. In India, more than 12 million biogas plants have been installed with a high rate of continued functionality. Since 1992, more than 140,000 biogas plants have been installed in Nepal, which has plans to increase availability of high-altitude digesters.

Page 79: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

79

Biogas production has improved livelihoods and decreased the strain on scarce resources by reducing the dependence on firewood collection. Through the displacement of firewood, indoor air quality has improved, positively affecting the health of women and children. In addition, the time savings for women are significant. In Nepal, an average of three hours a day can be saved by reducing the dependency on dispersed forms of biomass. Another livelihood benefit is the production of organic fertilizer—a main by-product of the anaerobic digestion process. The availability of the fertilizer saves farmers money and also reduces nutrient loading on fields. Finally, the construction and maintenance of biogas plants creates additional employment opportunities. In Nepal, it is estimated that 11,000 such jobs have been created.

The Netherlands Development Organisation (SNV) has committed to scaling up and expanding biogas development in Asia and plans to reach 1.2 million people with 210,000 additional biogas plants. Local financing issues, however, have proved significant barriers in Bangladesh and Cambodia. Carbon finance could prove to be one option for securing necessary funds for continued biogas development. Currently, the CDM has approved projects only for large-scale pig and dairy farms. Small-scale biogas programs are not eligible owing to difficulties in measurement, reporting, and verification (MRV). If the high transaction costs for small-scale projects can be eliminated, carbon markets could be a significant source of financing. Source: van Nes, W. J. 2006. Asia hits the gas: Biogas from anaerobic digestion rolls out across Asia. Renewable Energy World (January–February): 102–11.

4.6 Conclusions

Some conditions need to be met for realizing mitigation potential. The agricultural sector in Asia can play a significant role in GHG mitigation, but incentives to date are not conducive to investing in mitigation. At the same time, aligning growing demand for agricultural products with sustainable and emissions-saving development paths will prove challenging. Moreover, the carbon market for the agricultural sector is underdeveloped. To be sure, the verification, monitoring, and transaction costs are rather high, but the carbon market could be stimulated through different rules of access and operational rules in carbon trading, as well as capacity building and advances in measurement and monitoring. Finally, policies focused on mitigating GHG emissions, if carefully designed, can help create a new development strategy that encourages the creation of more valuable pro-poor investments by increasing the profitability of environmentally sustainable practices.

Key GHG low- or no-cost mitigation activities in the Asia and Pacific region include low or no till and other sequestration methods, as well as reducing methane emissions from rice fields. At a price of US$20/tCO2-eq, benefit streams from mitigation could add up to US$5.5 billion per year. Compared with total global economic mitigation potential Asia could mitigate approximately 18 percent of emissions at these carbon prices.

Page 80: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

80

CHAPTER 5. POLICIES, INVESTMENTS, AND INSTITUTIONS FOR ENHANCED RESILIENCE OF DEVELOPING ECONOMIES IN ASIA AND

PACIFIC IN THE FACE OF CLIMATE CHANGE

Key messages

Adaptation to climate change in the Asia-Pacific region will require advancements in three areas: (1) pursuing innovative adaptation to climate change, (2) strengthening important ongoing development initiatives to support climate change adaptation, and (3) implementing climate change adaptation investments and policies.

Levels and targeting of rural infrastructure investment will need to account for climate change.

Given the shifts in volume of rainfall and increased temperatures for the Asia Pacific region, as well as sea-level water rise and glacier melt, investments focusing on enhanced water control, water management and efficiency will be crucial for adaptation to climate change.

Knowledge and information sharing among farmers, government implementing agencies, and researchers should be given an enabling environment that supports adaptive management.

Better risk-sharing policies, likely provided by both the government and markets, such as weather-based crop insurance, need to be tested and implemented. An appropriate balance between public sector efforts and incentives, such as capacity building, the creation of risk insurance, and private investment, needs to be struck so that the burden can shift away from poor producers.

Crop breeding will be an essential component of adapting to key biotic and abiotic stresses related to climate change, including drought, heat, salinity, pests, and disease. Biotechnology and genetic modification will be an increasingly large component of crop breeding because of the nature of upcoming climate change stresses.

Given that international agricultural trade is an important mechanism for sharing climate change risk, open trading regimes should be supported.

Climate change can become the stimulus for implementing difficult but necessary changes. Rising prices of carbon, food, fuel, and environmental resources due to climate change could stimulate significant policy and investment opportunities.

Market-based approaches for managing environmental services in response to climate change (such as water pricing, payment for environmental services, and carbon trading) will be increasingly important.

Improved definition and protection of land and water property rights will be necessary to effectively implement market-based approaches to climate change policy.

Women’s roles in household and agriculture production, as well as their rights and control of assets, must be strengthened to avoid greater vulnerability under climate change.

Page 81: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

81

Asian countries should integrate climate policies and investments fully into development policies, while still recognizing the urgency of climate change adaptation. The need for additional funding also must be addressed; without this integration, climate adaptation plans may simply add another layer of planning rather than aiding the mainstreaming process.

The outcome of negotiations on a new international architecture for climate change policy will have profound implications for development financing for adaptation in Asian agriculture. Therefore, stakeholders need to emphasize the importance of adaptation and the synergies with mitigation in the agricultural sector in their recommendations and negotiations.

Effective cooperation among governments in Asia and the Pacific is necessary to ensure effective implementation of adaptation and mitigation strategies in their respective countries, as well as to explore financial means to address climate change.

Adapting farm management practices in Asian agriculture can have important synergy effects with mitigation (see measures outlined in Table 5.1).

5.1 Introduction

The current scientific consensus holds that greenhouse gas emissions and atmospheric concentrations are set to increase for some decades to come and that global mean surface temperature (and hence climate change with the impacts described in Chapter 2) will continue to increase long after the peak of emissions is passed. Even with an aggressive mitigation strategy, global surface warming will continue up to and beyond the end of the century. There is room for debate and uncertainty about how much warming there will be and at what rate it will unfold, but there is no doubt about the general trend of the curve. To maintain their present levels of prosperity and continue to develop, all countries have no alternative but to adapt to this change.

In the face of this adaptation imperative and current insufficient capacity to adapt, this chapter sets out to answer the following question:

Given the likely effects of climate change, the varied economies in developing countries in Asia and the Pacific, and their highly complex and dynamic socioeconomic and political environments, what initiatives to build resiliency and promote adaptation should different development actors implement to help achieve the Millennium Development Goals on poverty and hunger?

Achieving enhanced resilience in the face of climate change will require strengthening the adaptive capacity of countries in the Asia and Pacific region as well as implementing appropriate adaptation investments, policies, and institutions. Moreover, mitigation measures can support adaptation options and provide much-needed funds for further adaptation (Bryan et al. 2008; FAO 2009). Adaptation measures should be targeted to the countries, sectors, and people most vulnerable to the adverse impacts of climate change—that is, those most exposed to and most sensitive to the adverse impacts of climate change and those with the least adaptive capacity (Figure 1.1).

Both mitigation and adaptation response options need to be implemented by a variety of actors at local, community, national, and global levels. To decide which actor is most

Page 82: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

82

appropriate, it is useful to look at the time and spatial dimensions involved in the specific adaptation (and mitigation) response. Figure 5.1 provides examples of several common response strategies related to the agricultural sector. The spatial scale can help to identify what types of institutions are required, both for policies to set the enabling conditions and for action to carry out the necessary activities. These actions can take place at the global, national, local, or even individual level. Actions at the individual level, such as changing a crop variety or building a farm pond, generally do not require much in the way of institutions for coordination, although coordination at higher levels may be needed to produce new varieties and develop seed systems that distribute them.

Moving up the scale to response options that operate at the group or community level, such as a pond or small reservoir to serve a group or community, some form of coordination becomes necessary. At the local level, collective action institutions are often most appropriate, although some state institutions may also be relevant—for example, to provide technical advice to a group of farmers constructing or operating the reservoir. Moving upward on the spatial scale, local government or other state agencies become increasingly important for coordination, although collective action institutions may still be relevant, as in Nepal’s national federation of forest user groups.

The relative roles of state and collective action are illustrated by the triangles on the right hand side of Figure 5.1. In general, if the relevant scale for policies or actions is global, then international institutions are required—either existing ones, such as the UN agencies, or new institutions, such as carbon credit exchanges formed after the Kyoto Protocol in 1997.

The time frame for actions also provides insight into the nature of institutional arrangements needed. Although climate change response schemes arguably need to be set in motion very soon, some will show results in the short term (in a year or two), others over the medium term (in 2–10 years), and yet others over a much longer time horizon. The longer the time span between actions taken and results seen, the more difficult it will be to gain and maintain support and to monitor progress. Some actions, such as responses to crises like drought or flooding, will be intermittent. These actions call for institutional structures for preparedness and rapid response. The temporal scale may also indicate the relevance of property rights issues when there is a significant lag between an action and its consequences, especially between investment and returns such as for planting trees (Meinzen-Dick and Moore 2009).

At the center of agricultural adaptation are innovative responses to climate change, which are already in development but have not been implemented on a wide scale. These responses include changes in agricultural practices for crop and livestock systems. Enhancing the ability of farmers to respond to climate variability and climate change will require significant improvements in developing and disseminating agricultural technologies targeted to the major evolving biotic and abiotic stresses generated by climate change. Improved crop varieties have the potential to be more drought tolerant, increase nutrient-use and water-use efficiency, and decrease pesticide use. But new technologies, by themselves, are not sufficient to successfully address the challenges climate change poses for agriculture, including increased risks to production and household income.

To protect against the devastating outcomes from agricultural failures due to weather and climate, and to reduce risk aversion in farmer production decisions and thus enhance the potential for adoption of adaptive farming systems, programs and policies should be

Page 83: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

83

implemented to improve risk management and crop insurance, including climate-based insurance.

A stable and supportive policy environment that makes those programs available and profitable is also a critical factor. Such a policy environment requires strengthening important ongoing development initiatives to support climate change adaptation, which have been implemented to varying degrees throughout the developing world. These initiatives include secure property rights; improved economic incentives and green markets; improved information collection, use, and dissemination; extension services; and enhanced social protection and fiscal resiliency.

Finally, effective implementation of an agenda for climate change adaptation will require mainstreaming climate change and adaptation into development planning, reforming climate-related governance and institutions, and undertaking massive new investments. Thus, adaptation to climate change in the Asia-Pacific region will require advancements in three areas: (1) pursuing innovative adaptation to climate change, (2) strengthening important ongoing development initiatives to support climate change adaptation, and (3) implementing climate change adaptation investments and policies. These areas will be described in detail in the reminder of this chapter. Table 5.1 presents activities for areas (1) and (2).

5.2 Agricultural adaptation for the Asia and Pacific region

Adaptation to climate change is an adjustment made to a human, ecological, or physical system in response to a perceived vulnerability (Adger, Arnell, and Tompkins 2005; Figure 1.1), which is defined—according to the IPCC—as “the degree to which a system is susceptible to, or unable to cope with, adverse effects of climate change, including climate variability and extremes. Vulnerability is a function of the character, magnitude, and rate of climate variation to which a system is exposed, its sensitivity, and its adaptive capacity” (McCarthy et al. 2001).

As shown by the vulnerability indicators developed for this study, most developing member countries are considered moderately to highly vulnerable to climate change, with a low capacity to adapt to global warming as a result of low incomes and high poverty levels (Table 3.3 and Figure 3.4). A poor resource base, inequalities in income, weak institutions, and limited technologies further contribute to low adaptive capacity. Countries in particular need of adaptation strategies include Afghanistan, Bangladesh, Cambodia, India, Lao PDR, Myanmar and Nepal.

Decisions about which adaptation measures to adopt are not made in isolation by rural and agricultural individuals, households, or communities, but in the context of the wider society and political economy (Burton and Lim 2005). The choices are thus shaped by public policy, which can be supportive or can at times provide barriers or disincentives to adaptation. Possible supporting policies to stimulate adaptation measures are shown in Table 5.2.

Adaptation policy should be an extension of development policy that seeks to eradicate the structural causes of poverty and food insecurity. The complementarities between the two enable a streamlined approach toward achieving both adaptation and poverty alleviation goals. General policies that should be supported include promoting growth and economic diversification, strengthening institutions, protecting natural resources, creating markets in water and environmental services, improving the international trade system, enhancing resilience to

Page 84: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

84

disasters and improving disaster management, promoting risk sharing such as social safety nets and weather insurance, and investing in research and development, education, and health.

Adaptation options and their supporting policies should be adopted by the appropriate level of government and implemented by institutions in direct contact with beneficiaries. For example, adaptation responses such as changing planting dates and tillage practices may require technical services provided by local extension agents and coordinated by regional universities and research institutions. Agricultural research, including crop breeding to develop drought- and heat-tolerant crop varieties, will require both public and private investment. Structural adaptation measures, such as creating water markets and price incentives, will need to be implemented on a national level, most likely in partnership with organizations of users.

The challenge facing the global poor and those who would provide assistance is not solely a matter of finding ways of improving adaptation to anthropogenic climate change. Many communities, not necessarily limited to the poor, are not even well adapted to their current climate. The losses from floods, droughts, coastal storms, and other impacts are already unacceptably large and are increasing. These impacts can be attributed to anthropogenic climate change only to a relatively minor degree. People are now suffering, and economic development is being impeded by climate variability and extremes because their level of adaptation is below what it could be given the “availability” of adaptation measures. There is, quite simply, an adaptation deficit in relation to the existing climate (Burton 2004). It follows that any effort to improve current adaptation and adapt to anticipated future climate change has to be built upon the present circumstances and state of vulnerability. This challenge thus includes reducing the adaptation deficit even as we proceed to address adaptation to future and growing risks (Burton 2006a, b).

How can the adaptation deficit be closed and then adaptation policies and strategies extended to meet the challenges of climate change? Adaptation responses can be categorized by the level of ownership of the adaptation measure or strategy. Individual-level or autonomous adaptations are those that take place—invariably as a reactive response (after initial impacts are manifest) to climatic stimuli—without the directed intervention of a public agency and assuming efficient markets (Smit and Pilifosova 2001; Leary 1999; Mendelsohn 2006). Policy-driven or planned adaptation is the result of a deliberate policy decision by a public agency, based on an awareness that conditions are about to change or have changed and that action is required to minimize losses or benefit from opportunities (Pittock and Jones 2000). Thus, autonomous and policy-driven adaptation largely correspond with private and public adaptation, respectively (Smit and Pilifosova 2001). Table 5.3 provides examples of autonomous and policy-driven adaptation strategies for agriculture.

Autonomous adaptation responses will be evaluated by individual farmers in terms of costs and benefits. It is argued that farmers will adapt “efficiently” and that markets alone can encourage efficient adaptation in traded agricultural goods (Mendelsohn 2006). Yet in situations where market imperfections exist, such as in the absence of information about climate change and presence of environmental externalities and land tenure insecurity climate change will further reduce the capacity of individual farmers to manage risk effectively. Individual-level responses tend to be costly to poor producers and often create excessive burdens. As a result, an appropriate balance between public sector efforts and incentives, such as capacity building, the creation of risk insurance, and private investment, needs to be struck so that the burden can shift away from poor producers.

Page 85: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

85

5.2.1 Local coping strategies

Despite the commonalities of the natural disasters occurring due to climatic variability in the various regions of Asia, the coping strategies applied at the regional level vary. The variation in strategies may be due to geographical differences, social acceptability, farmers’ capacity (knowledge, materials needed), and availability of government support (such as technical and financial). These local coping strategies presented here by region will be beneficial for countries within the same region as well as those outside the region experiencing the same natural disturbances.

Central Asia

Mountainous areas in Central Asia experience extreme coldness that affects crop production—a condition also experienced in the Himalayas of South Asia and similar zones of East Asia. As a response to the extreme cold, Tajikistan farmers practice an alternative cultivation method that involves using cold frames to allow earlier seeding of plants (UNFCCC 2008a). This practice ensures continuous production of key crops despite extreme weather events and thus assures farmers’ income or even increases the potential for higher income. Another coping strategy applied by households in Tajikistan and normally done by women is a food preservation technique (UNFCCC 2008b). Before the onset of cold weather, women cure raw vegetables and can them to ensure available food for the family during winter. This option is equally relevant in Nepal, where processing of green leafy vegetables is done as postharvest management during extremely cold conditions and promotes local enterprise for mountain women (Manandhar 1998).

When extreme coldness is coupled with flooding, more disasters take place, such as land degradation, soil erosion, or deforestation. A comprehensive adaptation strategy is necessary to combat these natural disasters. Both the government and the affected communities need to work together and develop joint adaptation projects. Managing disaster risk through community-based management projects was found to be effective in Viet Nam (Francisco 2008) and thus may be applied to countries in Central Asia as well. Aside from the solidarity it creates within the community and with the government, this approach also disseminates knowledge and education on how to cope with these natural disasters, thus reducing risks in communities and saving lives. Other local coping measures adopted by communities in Tajikistan that can be applied in similar areas or countries in Central Asia are presented in Annex Table 5.1.

East Asia

In the loess highlands in western and northern China, farmers control soil erosion through a series of dams or dam-fields (UNFCCC 2008c). The dams control floods and retain water, while the dam-fields are used to receive mud flows from erosion and thus create new land for cultivation. This strategy, however, has potential mal-adaptation effects, including the inability to control soil erosion in the whole watershed, particularly at the sides and top of the hills, and salinization of the dam-fields (UNFCCC 2008c).

In the Tibetan Plateau, extreme cold reduces the survival or productivity of livestock. In western Sichuan in Southwestern China, livestock breeders select breeding jiulong (valley-type) and maiwa (plateau-type) yak during extreme cold weather (Wu 1998). This strategy ensures continuous production of yak and thus provides a source of food and income for farmers. It

Page 86: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

86

might be practical for other Asian farmers to check the feasibility of applying livestock breeding in areas affected by extreme coldness, or extreme heat, for that matter.

South Asia

Changes in climatic conditions result in flooding, erratic rainfall, drought or aridity, and sea-level rise in the countries of South Asia. In Bangladesh, there are two types of flooding—barsha, or moderate flooding that brings silt to agricultural land and thus increases soil fertility, and bonna, or high-intensity flooding (UNFCCC 2008d) that causes damage to agricultural crops, low survival or productivity of livestock, waterlogging, loss of livelihoods, and, in extreme cases, destruction of settlements and loss of lives. Farmers have come up with a number of coping strategies at the farm level as a means of survival during the bonna floods as well as waterlogged areas. Farmers in Jamalpur District and other coastal areas such as the Brahmaputra/Indo-Gangetic River Basin have established community rice-fish farms—integrated agriculture-aquaculture (IAA)—in floodplains or during the flood season (Dey and Prein 2005; FAO 2001). This system ensures food and nutrition availability, increases income, improves use of resources, and promotes community cooperation.

Another adaptation strategy common to most South Asian countries is appropriate crop selection as a response to flooding. To avoid the impact of floods, farmers in Bangladesh adjust their transplanting of aman (a wet season rice variety) (UNFCCC 2008e). The farmers plant early or late varieties of transplanted aman to avoid crop loss due to variations in flood recurrence. The early production of rice encourages the growing of other additional crops. This practice enhances incomes not only from rice production, but from other crops as well. Farmers in Uttar Pradesh, India, may benefit from such a coping strategy given that flooding in this area is similar to that in Bangladesh.

Hydroponics is another method of cultivating crops during the flood season, especially in waterlogged areas (UNFCCC 2008f). Crops, mostly vegetables, are grown in floating gardens. This practice ensures subsistence food during flooding and may be a potential source of additional income. Duck raising as part of livestock production may also be exploited during the monsoon period. Mallick (2006) explains that raising ducks and diversifying the diets of communities are coping strategies during the flood season in Bangladesh. Hydroponic vegetable farming might be an option for the Mekong Delta as well.

Rising sea level results in flooding leading to waterlogged areas. In Goa, India, farmers practice khazan—a traditionally community-managed IAA system—in waterlogged areas. Aside from establishing cooperation within the community, IAA promotes a mutually beneficial relationship between the rich and the poor through employment generation and labor sharing (TERI n.d.).

Another natural disaster of significance in South Asia is drought or aridity. In general, the most common adaptation strategy consists of sustainable water management through tanks and dams. In Sindh, Pakistan, laths (temporary structure of 1-3 m high used as traditional spate [flood] irrigation) are used (UNFCCC 2008g). In India, anicuts (small to medium-sized dams) are used to harvest rainwater and serve as reservoirs (Narain, Khan, and Singh 2005). Other rainwater-harvesting techniques include underground tanks or kunds in the Thar Desert, India (UNFCCC 2008h); gutters and pipes to collect rooftop rainwater in Bangladesh (UNEP DTIE 2000); bamboo stems for drip irrigation in Bhutan (UNFCCC 2008i); ground barriers (contour

Page 87: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

87

bunds, nallan bunds, gabions) and shallow excavations (contour trenches, farm ponds, reservoirs in bedrock) in Maharashtra, India (Sivanappan 1997); and cascading tanks in Sri Lanka (Herath 2001).

Other coping strategies involve appropriate crop selection. In the Barind Tract15 of Bangladesh, farmers plant drought-resistant fruit trees such as mangoes or engage in jujube gardening (Selvaraju et al. 2006). Domesticating indigenous varieties of cereals and fruit trees promotes local enterprises for women in Arunachal Pradesh, India (UNFCCC 2008j). Alternative cultivation methods such as seedbed methods for transplanting seedlings (UNEP DTIE 2000); home gardening (UNFCCC 2008k); and rotational cropping (Verma 1998) are also helpful in increasing crop production as well as ensuring food availability during adverse climatic conditions.

Erratic rain can result in soil erosion and land degradation. Methods of controlling soil erosion in the Himalayas include terracing, field leveling, ploughing, sheet erosion control, and biofencing (Verma 1998). Application of manure or ash from organic manure, crop residue, or kitchen ash can enhance soil fertility (Verma 1998).

Farmers in the western Himalayas rely on meteorological indicators and animal behavior to predict rainfall (UNFCCC 2008l). Indicators include the visible spectrum around the sun and the moon; clouds or wind direction; the activities of various birds, animals, and insects; and condensation.

Southeast Asia and the Mekong region

As in the IAA system in South Asia, farmers in West Java, Indonesia, grow fish in huma or dry swidden fields during drought conditions and in sawah or wet fields during flooding (FAO 2001). This alternative cultivation method encourages generation of cash income and food availability for farmers during extreme weather conditions like drought and flood. During drought conditions, farmers in the Philippines are encouraged to (1) change cropping schedules to lessen demand for irrigation or adjust the cropping calendar according to water availability; (2) line canals to reduce water losses; (3) maximize the use of available water during abundant periods by constructing reservoir-type projects; (4) redesign irrigation facilities to reuse return flows; and (5) introduce other water-saving techniques (Lansigan 2003). Some traditional farming practices like drip irrigation, mulching, and other improved irrigation methods, as well as windbreaks to minimize wind speed and evapotranspiration, can improve the use of dwindling irrigation water (Baradas and Mina 1996 in Jose and Cruz 1999). Drought-tolerant crop varieties and efficient farming practices should likewise be considered. Boer (2009) presented the same strategic options found in the Philippines for farmers in Indonesia. Aside from improved crop technologies and water efficiency, Boer suggested the creation of climate field schools (CFSs) to develop farmers’ capacity in climate forecasting information and risk management. Boer (2009) further clarified that CFSs go beyond farm level. Off-farm programs on agribusiness can help farmers estimate production periods for agricultural commodities based on climate forecasts and thus can help them take advantage of expected price changes for these commodities. Furthermore, such programs can increase farmers’ bargaining power through enhanced collaboration with the government, the private sector, farmers’ organizations, and other groups.

15 The Barind Tract includes Dinajpur, Rangpur, Pabna, Rajshahi, Bogra, and Joypurhat of Rajshahi Division, Bangladesh.

Page 88: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

88

In the Mekong Delta, communities in Attapeu province, Lao PDR, diversify their diets during the flood season from rice-based diets to edible aquatic resources like fish, crabs, and other food from the Delta. Prolonged food shortages, however, threaten wetland and forest resources in the Delta (Meusch et al. 2003). Viet Nam illustrated a successful community-based adaptation strategy as response to climatic changes in Quang Dien and Phu Vang Districts, Thua Thien Hue province, in the north-central coast of the country (Francisco 2008). Affected communities and the government worked together to build capacity for adaptation to climate change. The critical steps in this effort were scenario building, planning, and implementation of projects. The main objectives were to help build communities’ adaptive strategies in the face of recurrent climatic catastrophes and to minimize loss of lives and property (Francisco 2008).

Details of local coping strategies in other Asian countries can be found Annex 5.1.

5.2.2 Innovative adaptation to climate change

Changes in agricultural practices

In agriculture, forestry, livestock operations, water resources management, public health, and other fields affected by climate change, multiple adaptation measures may typically be taken. Key changes in farm management practices include land-use changes to maximize yields under new conditions; application of new technologies and changes in input use; new land management techniques; changes in crop and livestock varieties; changes in planting dates; and water-use efficiency techniques. Changes in agricultural water management practices are discussed in a separate section. Adaptive agricultural management practices include effective use of pest, disease, and weed management systems through wider application of integrated pest and pathogen management techniques and development and use of crop varieties resistant to pests and diseases, as well as efficient quarantine capabilities and monitoring programs (Box 5.1). Changes in location or timing of cropping activities must be considered. Matthews et al. (1997) found that changing the planting time can lessen the negative impacts of extreme temperatures. Farmers in the Mekong Delta of Viet Nam are using a shorter-cycle rice seed variety to adapt to climate risks (Oxfam 2008).

Box 5.1. Zero tillage: An effective mitigation and adaptation strategy in South Asia

In the past decade, farmers in the rice-wheat farming system in the Indo-Gangetic plain of Bangladesh, India, and Pakistan have widely adopted minimum-tillage practices, which conserve resources under climate change. Since being introduced by researchers from a consortium of international agricultural research centers (IARCs) and national agricultural research systems (NARSs) in the late 1990s, zero tillage for wheat has been adopted rapidly, reaching to more than 1 million farmers on an estimated 5.6 million hectares (Rice-Wheat Consortium 2005). Such rapid and widespread adoption of a natural resource management innovation is rare, although zero or minimum tillage has been adopted on a large scale in intensive mechanized farming systems elsewhere, with global adoption estimated to be as high as 90 million hectares (Murray 2005). Farmers’ wheat yields have reportedly improved, and production costs have decreased by an average of US$65 per hectare, with additional benefits for water conservation and herbicide reductions (Hobbs 2001).

Field-based livestock systems require extra attention in implementing adaptation measures (Howden et al. 2007), including matching stock rates with pasture production, variable rotation of pastures, and changing grazing times and production periods. Additional measures

Page 89: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

89

include shifting integration of mixed livestock-crop systems using adapted forage crops, reassessing fertilizer applications, ensuring plentiful water supplies, and utilizing supplementary feeds and concentrates (Batima et al. 2005; Adger et al. 2003; Daepp, Nosberger, and Luscher 2001).

Changes in climatic conditions influence intensive livestock production. During warm weather, there is less need for winter housing and feed concentrates. On the other hand, warm weather requires increased management and infrastructure to lessen the detrimental effects of heat-related stress on productivity, fertility, and fatality (Howden et al. 2007). Heat-tolerant livestock breeds have lower levels of productivity (Howden et al. 2007), suggesting a need for additional research in higher-yielding heat-tolerant breeds.

In any given situation or context, the choice of adaptation measures may be difficult and constrained by expense, by lack of knowledge on how to implement them, and by traditional beliefs and cultural practices. Notwithstanding these impediments, farmers and others at risk from climate change can receive several forms of external help. Possibilities include technical information and advice or guidance; weather and seasonal climate forecasts and warnings; drought or flood relief; and insurance or other forms of financial assistance and risk spreading. These actions can be taken to reduce exposure or vulnerability to risk where the poor agriculture- or resource-dependent population lives. Poor farmers are not passive recipients of external assistance. They can and do take other initiatives, such as diversifying their sources of income by beginning other enterprises at the village level or by migrating temporarily or permanently to towns or cities in search of other kinds of employment.

A combination of these suggested adaptation measures for cropping systems will have substantial potential to reduce the destructive effects of climate change in agriculture. Other types of support are also required, however, because farmers cannot on their own adequately adapt to climate change and variability (Box 5.2).

Box 5.2. Coping versus adaptation: Examples from South Asia

Smallholder farmers in the semi-arid Jhalawar district in Rajasthan, India, are highly vulnerable to climate variability, such as consecutive droughts. In 2002, Jhalawar experienced its fourth consecutive year of drought. To better cope with climate variability, farmers have shifted from traditional crops, such as sorghum and pearl millet, to soybean, which receives higher market prices and yields quick returns because of its shorter life cycle (Kelkar and Bhadwal 2007). In the Lakhakheri Umat village (Jhalawar, Rajasthan, India; TERI 2003), where nearly all of the farmers have small or marginal landholdings, farmers use a variety of coping mechanisms, such as selling cattle, shifting to other crops and labor, and migrating seasonally.

In addition, women farmers in Bangladesh in flood-prone areas are building “floating gardens” made of hyacinth rafts to grow vegetables during the flooding season (HDR 2008). These options are temporary coping measures, however, that do not prepare farmers for future climate problems. As a result of lack of awareness, procedural complexities, and stringent eligibility criteria, farmers do not use options that improve long-term adaptive capacity such as institutional credit, crop insurance, and drought-resistant varieties (Kelkar and Bhadwal 2007).

Muhammed et al. (2004) reports on coping strategies practiced in vulnerable regions of South Asia. During drought periods, farmers in India and Pakistan borrow money from lenders and banks and some migrate to search for alternative livelihoods. Other adaptation options include buying or saving fodder for livestock given changing feeding patterns of livestock, selling livestock and other belongings,

Page 90: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

90

shifting livestock to other areas, planting less water-intensive crops, selling or mortgaging property, and if available, working in government-sponsored food- or cash-for-work programs.

Furthermore, farmers in Bangladesh, India, and Nepal cope with flooding by migrating to look for alternative livelihoods, engaging in off-farm activities, protecting livestock, applying for insurance for local crop varieties, harvesting and trading premature fish to avoid escape and loss, spending savings, and securing loans from the informal sector.

To move beyond short-term coping, effective farm-level adaptation requires access to improved agricultural technology. Additional information on coping strategies and possibilities for future adaptation are presented in Annex 5.1.

Changes in agricultural water management

Water management adaptation measures that are being applied by farmers include wider application or use of practical technologies such as water harvesting, soil moisture conservation techniques (such as crop residue retention), and effective use and transport of water during drought periods (Howden et al. 2007). Some of these water management practices can prevent waterlogging, erosion, and nutrient leaching when rainfall increases.

Climate variability is increasing in places where it is already highest. The reduced storage of precipitation as snow, as well as the earlier melting of winter snow, is leading to shifts in peak runoff away from the summer season when demand is high in parts of Asia and elsewhere. Low-lying coastal areas affected by sea-level rise are experiencing inundation and increased damage, with storm surges and increased saline intrusion into vulnerable freshwater aquifers. And nonrenewable groundwater resources are being depleted. Consequently, increased flexibility in infrastructure and operations of irrigation systems—particularly large irrigation systems—will be crucial. Water-delivery systems need to be flexible (technically and institutionally) to deliver water for multiple uses (agriculture, environment, city, industry, and energy generation), from entire river basins down to and within large irrigation systems and under new ranges of water availability. The modernization of irrigation systems, in particular the establishment of better control systems at key distribution points, can increase farmers’ access and control over irrigation water resources, conserving water resources and enhancing adaptability to climate change (Renault, Facon, and Wahaj 2007). This improvement will be particularly important for the large surface irrigation systems fed by glaciers and snowmelt in China and India, but also for the large systems found in much of Central Asia.

In fragile upstream watersheds, where a combination of irrigated agriculture, rainfed agriculture, pasture, and forestry is practiced, a holistic approach to watershed management will be more important to adapt to more erratic rainfall events. This concern will be important for Lao PDR, Nepal, and Viet Nam, as well as some of the island states.

Water storage will be a key adaptation strategy, in the form of seasonal storage systems in the monsoon regions, where peak flood flows are likely to increase. Water storage comprises much more, however, including a continuum of surface and subsurface water storage options ranging from natural wetlands and water stored in situ in the soil through to rainwater-harvesting ponds and small and large reservoirs. Concerns about the negative social and environmental impacts led to reduced investment in large dams in the 1990s. Now, however, given the need to produce more food, provide stable water supplies for growing urban areas, and provide more energy resources, investments in large dams in Asia are increasing again.

Page 91: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

91

Investments in supplemental irrigation will be important to reduce the consequences of irregular rainfall through short-term interventions to capture and store more soil moisture or runoff. This approach will be particularly important in the semi-arid and arid areas of Central Asia, Afghanistan, parts of India, and some of the Pacifc Island states.

Large-scale groundwater development in Asia was developed in response to availability of cheap pumps from China and unreliable or unavailable access to surface water sources, particularly in India and parts of Southeast and East Asia. Groundwater now accounts for 50 percent of irrigation supply in South Asia and perhaps two-thirds of supply in the grain belts of northern China (Giordano and Villholth 2007). Groundwater development can be an effective adaptation to climate change, given the just-in-time availability and high efficiency of use of the resource. Some of the adaptation benefits of groundwater irrigation may be offset, however, by carbon dioxide from energy used to deliver the water or from N2O emissions from higher moisture. Moreover, in coastal areas, groundwater will be affected by saline intrusion as a result of sea-level rise. Conjunctive surface and groundwater management and economic incentives for reducing unsustainable groundwater use are important avenues to sustainably continue groundwater use in India, China, and other parts of Southeast Asia.

One avenue for both adaptation and mitigation might be treadle pump development. The Energy and Resources Institute (TERI) estimated for International Development Enterprises in India that that the operation of one treadle pump annually reduces CO2 emissions by 477 kilograms (kg) (TERI 2007a). The total emission reduction was quantified at 150,000 tCO2-eq for treadle pumps sold between April 2001 and March 2004. The entire project is estimated to generate reductions of more than 800,000 tCO2e in its lifetime. Given that water scarcity is expected to increase in parts of Asia as a result of global warming and other drivers, application of water-conserving irrigation technologies will be an important adaptation strategy. TERI (2007b) found that for four study regions in India, micro-drip irrigation saved an average of 54 percent of water resources and 39 percent of electricity compared with flood irrigation. This result is equivalent to an average annual CO2 emission abatement for every acre of drip adoption of 675 kg per acre per year.

Agricultural diversification

Many adaptation strategies are forms of agricultural diversification, including some of the farm-level adaptation strategies described in Box 5.2. One example of successful farm-level diversification has been alternative rice-shrimp farming in the Mekong Delta of Viet Nam, facilitated by flexible water control structures that allow for both freshwater and brackish water control.

Diversification into off-farm employment and seasonal migration are strategies that have been adopted in many Asia-Pacific countries as a result of resource scarcity, particularly small farm sizes and lack of income opportunities in farming for many years. In Indonesia, 34 percent of rural employment was in the nonfarm sector and nonfarm income provided 43 percent of total rural income in 2002 (SEARCA and IFPRI 2004).

Agricultural science and technology development

Technological change that increases agricultural productivity growth saves land and water, and increases the flexibility of cropping systems is essential if the challenges of climate change for the agricultural sector are to be met. In addition to conventional breeding, biotechnology and

Page 92: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

92

genetically modified (GM) crops are also likely to be essential tools for adapting to increased climate stress. They have the potential to increase crop adaptation to heat and drought stress while improving crop productivity, mitigating greenhouse emissions from fertilizer use, reducing pesticide and herbicide applications, and modifying plants for use as biofuel feedstock. Investments in biotechnology, including GM crops, could provide a transformational approach to addressing the trade-offs between energy efficiency and agricultural productivity.

Biotechnology tools—including DNA sequencers, chip-based gene expression, molecular markers, and many others—are revolutionizing crop improvement. Continued improvements in high throughput technology (a scientific experimentation method allowing scientists to quickly conduct millions of genetic and other types of tests) will make gene discovery for crop improvement routine and inexpensive. Complete or draft genome sequences for rice, poplar, grape, papaya, and maize are now available, and sequences for soybean, sorghum, and canola will be available in 2009, with complete genome sequences for all important crop species complete by 2015. Moreover, crop cultivars with GM traits have been broadly commercialized in the past 12 years. In 2007, transgenic varieties, most containing insect- and/or herbicide-resistance traits, were grown on 114.3 million hectares (ha), primarily in Argentina, Brazil, Canada, China, India, and the United States (James 2007). Many more crops and traits are currently in development and are slowly entering the regulatory pipeline (Atanassov et al. 2006). Farmers’ experience with GM crops has been largely positive, with increased management options, reduced pesticide use, and in some cases improved yields (Brookes and Barfoot 2007). This experience suggests that GM crops are becoming an established technology in these countries at the early stages of application.

Potential of improved crop varieties to increase nutrient use efficiency and decrease pesticide use. Climate change is projected to increase the pressure from both insects and weeds that can outcompete existing crop varieties. Breeding programs in developing countries are developing seeds that are high yielding in a given biotic and abiotic environment. Successful breeding programs have helped produce crop varieties that are resistant to a number of pests and diseases. Recent developments in GM research have produced soybean, rapeseed, cotton, and maize for herbicide tolerance, and other varieties are being developed to resist various pests and diseases (Phipps and Park 2002).

Pest-resistant and herbicide-tolerant technologies can potentially reduce pesticide and herbicide use, thus reducing harmful environmental impacts such as water pollution, while also improving yields. Their impacts are mixed, however. Insect-resistant GM crops—notably Bt cotton—reduced the amount of pesticide applications by 1.2 million kg between 1996 (when Bt cotton was introduced) and 1999. This amount is equivalent to 14 percent of all insecticides used (James 2000). But the impact of herbicide-tolerant crops on the total amount of herbicides used is ambiguous. Although herbicide-tolerant crops have reduced the number of active ingredients sprayed, they have not changed or have slightly increased the weight of the herbicides used (Benbrook 2001).

Synthetic nitrogen fertilizers aid crop growth but are also major contributors to GHG emissions. Nitrous oxide (N2O) is the fourth-largest contributor to the natural greenhouse effect. More efficient nitrogen use by crops has several important environmental advantages, in addition to lowering production costs for farmers in light of high fertilizer prices. Genes have been identified that improve the efficiency with which plants use nitrogen fertilizer, and GM plants

Page 93: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

93

with these genes are currently being characterized under field conditions. The reduced need for synthetic nitrogen fertilizers will reduce energy costs and help lower GHG production.

In the longer term, additional fundamental breakthroughs could be made. A crop’s ability to produce yields across many different growing environments is complex and can be affected by many different genes. The genes involved in determining yield potential and their importance and expression patterns vary widely depending on the crop and growing environment. Even so, genes affecting yield directly have been identified and are being evaluated in the field. Research continues to identify approaches that could reduce the inefficiencies of C3 photosynthesis or even convert C3 crops such as rice to more efficient C4 photosynthesis, which benefits from elevated atmosphere concentrations of CO2 (Normile 2006). C4 plants are better adapted to arid conditions because they have improved water use efficiency. The technical hurdles for this approach are high, but it is realistic to expect that these improved crops will be available in field trials within the next 10 years.

Biotechnology could profoundly affect future demand for freshwater and investment requirements in irrigation and other water sectors. GM crops have the potential to address major water-related stresses under both rainfed and irrigated farming and to possibly offer solutions to important water quality problems. Breeding crop varieties with high water-use efficiency—a good indicator of the crop’s ability to withstand environmental stresses, particularly drought and salinity—is one policy option. Many genes associated with adaptation to various types of stress tolerance have been identified and incorporated into crops. These stress-tolerance genes are being field tested in maize, rice, wheat, and soybeans and will be developed in other crops.

Plant breeding (conventional and molecular) has been and remains an important tool for dealing with drought and should continue. GM approaches appear to offer more genetic variations that could lead to further advances. Despite commercial selection of major crops for drought tolerance for the past 50 years, together with traditional selection for several preceding centuries, adequate water is still the factor that limits crop production more than other constraints. The use of breeding for improving drought tolerance has been well tested, and the degree of improvement is well understood. Initial experiments and field testing with transgenics suggest that higher levels of drought tolerance appear possible. Most interestingly, progress on drought tolerance may be possible without interfering with yields under good conditions, which is often a trade-off with conventional breeding.

Condon et al. (2004) discuss three main processes that crop breeding can use for high water-use efficiency: (1) moving more of the available water through the crop rather than letting it go to waste by evaporating from the soil surface, draining beyond the root zone, or remaining behind in the root zone during harvest; (2) acquiring more carbon (biomass) in exchange for water transpired by the crop (that is, improving crop transpiration efficiency); and (3) partitioning more of the acquired biomass into the harvested product. These processes are interdependent, and their relevance depends on water availability during the crop cycle. Because these crops are not yet on the market, crop simulation modeling can be used to assess the likely impact of changing the expression of crop traits on water-use efficiency and yield (Condon et al. 2004). Biotechnology’s role as a possible substitute for large-scale water investments must be considered in future planning for irrigation and water supply and sanitation investments.

Risks and limitations. In spite of the prospects for GM crops, biotechnology is controversial, and public acceptance and safety issues must be resolved (Rosegrant, Cline, and

Page 94: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

94

Valmonte-Santos 2006). Food safety risks are often raised, but no documented case of food safety problems or negative human health impacts from GM food crops has arisen, despite many years of experience. Frequently cited environmental risks include the possibility of outcrossing, such as the development of aggressive weeds or wild relatives with increased resistance to diseases, or the rapid creation of new pest biotypes adapted to genetically modified plants that could upset the ecosystem balance. In addition, crop biodiversity may decrease if a small number of genetically modified cultivars displace traditional cultivars (FAO 2000). The biodiversity issue for GM crops is no different than for modern varieties in general. Introducing GM crops will affect crop biodiversity only if the genetic trait is introduced in one variety and planted as monoculture. If, however, it is introduced in several varieties (as in India, where more than 110 varieties of Bt cotton are growing), it should not have any specific negative effect on crop biodiversity (Gruere et al. 2008).

Implementing a transformational technology. Advances in the biotechnology underlying GM crops provide the capability to modify crops more rapidly and with fewer unpredictable changes than conventional breeding, and considerable resources have been invested in building scientific capabilities for crop improvement. Confidence in the ability to evaluate the risk of GM crop varieties will be necessary to capture the benefits of these technical advances. A number of steps need to be taken to improve the adoption and benefits of biotechnology and GM crops.

To translate technical advances into products that can improve crop production under climate change, public and private sector organizations need to develop additional capacity to address complicated intellectual property, risk management, and regulatory requirements. Additionally, the emergence of private sector crop improvement has resulted in opportunities for the private and public sectors to work together, but only if there is suitable understanding of the concerns of both sectors.

Lack of access to intellectual property and freedom to operate represent hurdles for those seeking to obtain privately owned technologies. Companies have shown a willingness to license genes and other technologies to public sector organizations if they are assured that those using their technology will respect intellectual property rights and will not expose them to liabilities that they cannot anticipate and control. Understanding the changing landscape of intellectual property, coupled with ongoing efforts to license protected technologies, can increase the available approaches for crop improvement for the public good.

In many cases, public–private partnerships (PPPs) will constitute the best mechanism for ensuring broad access to improved cultivars by identifying and encouraging effective plant breeders’ rights, intellectual property regimes, and technology transfer mechanisms. Policies that support the development of PPPs will increase access to advanced crop improvement technologies where conditions are not yet adequate to promote private commercial seed companies. Specifically, improvements related to climate change—such as nitrogen-use efficiency and water-use efficiency—are critical for developing countries.

The potential importance of PPPs to agricultural biotechnology research is well recognized (see Spielman et al. 2007a, b; Spielman et al. 2006; Pingali and Traxler 2002; Pray 2001). Examples for the Asia and Pacific region are shown in Box 5.3.

Page 95: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

95

Box 5.3. Private-public partnerships for biotechnology development in the Asia-Pacific region

Bt cotton in India. At present, the development and diffusion of Bt cotton in India is being driven by a number of close public-private interactions. Key proponents include (1) private sector leaders in the crop-science industry, namely Monsanto and MAHYCO (based in Jalna, India), (2) public research institutes such as the National Botanical Research Institute in Lucknow and the Indian Institute of Technology in Kharagpur, and (3) domestic seed companies operating throughout India. Through a complex web of research joint ventures and licensing agreements, Bt cotton varieties are being adopted rapidly in India, providing small farmers with new choices and options.

Biofortified rice in Asia. A unique public-private partnership under the auspices of the Golden Rice Humanitarian Board is currently leading the research and development of high-beta-carotene rice in Asia. The board’s role has been to address the issues of intellectual property rights to enable royalty-free transfer and commercialization of the technology, and it has succeeded largely thanks to the direct involvement of Syngenta, a Swiss company that negotiated to secure access to key technologies used in the Golden Rice research. These negotiations have allowed for the issuance of royalty-free sublicenses to public research institutes in Bangladesh, China, India, and the Philippines so they can develop locally adapted rice varieties with high beta-carotene content (GRHB 2006).

Bt brinjal (eggplant) in India. A partnership that aims to make Bt technology in brinjal affordable to farmers in Asia and the Pacific has been developed recently between the public and private sectors. Under the Agricultural Biotechnology Support Project II, an initiative supported by the U.S. Agency for International Development (USAID), MAHYCO is providing the technology to public sector research institutions in Bangladesh, India, and the Philippines, which will use the MAHYCO material to backcross with their own brinjal varieties. No royalties are required to be paid so long as the public institutions are not involved in commercializing the Bt varieties, and farmers will be permitted to save seed to cultivate crops in subsequent seasons (Balaji 2006; CU 2005).

Policies that favor private sector investment in crop improvements targeted to climate change in the developed and developing world are critical. These policies include (1) decreasing the bureaucratic hurdles to business formation; (2) developing infrastructure that enables production and distribution of improved seeds and other agricultural inputs; (3) developing appropriate regulatory and biosafety protocols for introduction of transgenic cultivars; and (4) reforming intellectual property rights that could encourage private investment in crop improvement.

Developing countries have chronically underinvested in science, technology, and innovation (Pardey et al. 2006). In most of the developing world, the growth in public investments in research stagnated after the 1980s. Investments in biotechnology and biosafety, especially by the public sector, may be insufficient to address pressing needs in both areas, especially when focused on resolving national constraints. In spite of the limitations, the public sector in many developing countries has invested in agricultural biotechnology research (Atanassov et al. 2006), yet few of its technologies have made it to the commercialization stage (Cohen et al 2005). Many developing countries, particularly those in Sub-Saharan Africa and Southeast Asia, need to develop the minimal infrastructure and scientific capacity to master and implement risk assessments and biosafety regulations.

What is needed is not necessarily more biosafety regulation, but effective biosafety regulation. Additional regulations, unnecessary procedures, and regulatory time delays tend to increase the costs of developing GM crops and complying with biosafety regulations. Additional

Page 96: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

96

unnecessary costs reduce the present value of GM crops and may even prevent the release of the technology. In most cases, however, the present value is affected more by regulatory time delays than by increases in cost. Therefore, cost in the sense of both time and money, becomes a barrier to entry for private companies, and especially the public sector.

Broader adoption of well-established productivity improvements, including intensive cropping, more efficient irrigation, and increased use of inputs such as fertilizer and improved conventionally bred crop varieties, can and should have a significant effect on crop productivity and sustainability. The increasingly mature technologies underlying GM crops suggest, however, that they represent a transformational technology for responding to climate change.

Agricultural advisory services and information systems

Effective dissemination of modern technologies is responsible for a considerable share of the success in Asian agriculture. The performance of agricultural extension has declined significantly in the past two decades, mainly because of the prevalence of supply-driven public extension services characterized by weak human capacity, limited coverage, and poor financial resources. Involving producer organizations in extension activities helps engage producers in programs that coincide with their own goals. There is a growing consensus that a mature extension system is characterized by a pluralistic system of extension funding and service provision (see Box 5.4 on Indonesia’s extension system). Farmers could contribute to the cost of extension services, but there is concern that this step would limit access by small farmers. Hence, a number of studies have concluded that commercial farmers should pay for extension advice, and the government should provide complimentary extension services to small producers. The public sector must continue to be a major player, however, both in funding and in coordinating operations.

Extension policies and strategies need to define an effective division of labor between public extension and other providers and identify overall objectives for public sector involvement in extension. Another challenge to privatizing extension services is the lack of private providers, especially in remote areas. In countries that have privatized provision of advisory services, many service providers have emerged, with many nongovernmental organizations (NGOs), private companies, and semi-autonomous bodies delivering extension advice to farmers. The large number of service providers has led to the need for coordination and regulation, because different providers have offered conflicting technical recommendations in some cases. A pluralistic agricultural extension system also allows for complementarity of providers. Underscoring the importance of pluralism, one study showed that NGOs tended to promote natural resource conservation more than public advisory service providers.

Page 97: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

97

Box 5.4. Extension in Indonesia

Indonesia’s experience of decentralizing its extension system has been mixed. Sharp reductions in funding and the removal of centralized guidance have had adverse impacts on extension. There have also been successes, however, in the form of management experimentation, participatory approaches, dissemination of market and upstream information and technology, decentralized services, and some movement toward privatized extension. Indonesia can make use of several relevant avenues for developing extension services, including the following:

Expanding the coverage of the Decentralized Agriculture and Forestry Extension Project or similar agricultural extension projects. These projects were originally funded at the national level but are gradually being taken over by district governments. Such projects could provide necessary guidance and training, while demonstrating to district governments the importance of agricultural extension activities in improving farmer incomes.

Implementing farmer field schools using participatory methods to help farmers develop analytical skills, critical thinking, creativity, and decision making. Participatory extension, however, requires a simple curriculum, short-duration training, and high-quality trainers. Prospects for collective action to improve outcomes are greater when larger groups of farmers within a village are trained.

Privatizing parts of extension through contracting, for example, by seed companies. This approach can introduce incentives for higher efficiency. Success is increased when extension is linked to the delivery of a specific technology (such as hybrid maize or poultry) and to larger, more homogeneous groups of farmers. For commodities where private extension services cannot be self-supporting, the government needs to continue providing assistance and training.

Training field extension personnel in a broader range of subjects, not limited to technology. They should be provided with additional resources as needed to help them advise farmers on diverse issues such as how to obtain credit, add value to their agricultural products, and obtain markets for their products (SEARCA and IFPRI 2004).

Successful action in agricultural adaptation requires better and clearer information combined with investment and advisory services to disseminate the information to users, as well as feedback loops to generate bottom-up information from farmers, foresters, and fishers. Information is an important component in all successful management reforms. Improved information systems allow for more informed decisions, heightened awareness of the impacts of people’s actions, and greater incentives to change crops and adopt practices to enhance management sustainability. As a basis for adaptation planning, developing countries, alongside their international partners, will need to conduct comprehensive climate change monitoring and forecasting. In most cases, these activities will require developing countries to allocate more resource for collecting systematic meteorological data and developing stronger human capacity in climate change analysis and research. Until this capacity is developed, the international research community will remain critical to these efforts.

More advanced information technologies are developing quickly and will be increasingly important. Satellite remote sensing to measure water productivity and spatially disaggregated patterns of land use and geographic information systems have been successfully used and should be expanded dramatically in seeking land- and water-saving policies in response to climate change. Both policymakers and local communities require a combination of technical expertise and local knowledge. In many cases they will require more effective innovation systems that

Page 98: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

98

disseminate information about adaptive land and water management practices—both new technologies and practices developed by farmers—and about their consequences over both space and time. Participatory land use planning can build on technical models as well as on systems of problem identification, farmer field schools, and other methods to identify both the constraints and opportunities, especially in the context of climate change.

Risk management and crop insurance

Crop insurance has historically been relatively ineffective, even in developed countries, and problems are greater in developing countries. At present, communities and individuals in most developing countries lack insurance coverage against extreme climate-related weather events. Index-based insurance and credit may overcome some of the limitations of traditional agriculture insurance, allowing farmers to take the increased risks that tend to be associated with higher-yielding production decisions that can lead to increased incomes and agricultural productivity (Tubiello et al. 2008). Rather than basing indemnity payments on individual farm yields, index-based policies pay policyholders based on an index of, for example, area or regional yields or weather data such as temperature or rainfall. This approach reduces the transaction costs of monitoring involved in traditional insurance products. Because farmers are paid regardless of their individual yields, this approach also encourages farmers to continue producing if possible (Kryspin-Watson 2006).

The private sector is often reluctant to provide crop insurance because of high implementation costs and the fear of large losses in catastrophic events that are unlikely to be covered by income from insurance premiums. Public-private partnerships could overcome these limitations, thus serving three purposes. First, they could perform the classic insurance function of spreading risk. Second, they could ensure continuity of government operations after a severe loss event. Third and most important in the adaptation context, they could help to ensure that adequate adaptation measures are taken. Insurance in this case would be an instrument of public policy and not an end in itself. The objective would be to maximize agricultural productivity in the face of increased climate shocks. Insurance would encourage, facilitate, or even mandate adaptation measures. An innovative approach to a comprehensive insurance program would contribute to these goals. Insurance could be made available at concessionary rates (thus contributing to meeting the UNFCCC obligation to help developing countries meet the costs of adaptation), subject to the condition that the insured activity or the property meets certain adaptation or vulnerability reduction requirements.

Box 5.5 Weather-based insurance in India

In 2003, a pilot program for weather insurance was launched in the Andhra Pradesh province, India, to help protect famers against low rainfall. Implemented by BASIX, one of India’s largest microfinance institutions, the program began with 250 policies sold to groundnut and castor farmers in region. The index-based weather insurance relied upon rainfall data in the province and made payments to farmers when rainfall fell below a predetermined amount. Based on feedback from farmers, BASIX expanded the project in 2004, selling more than 700 policies. In 2006, BASIX sold rainfall and mixed weather contracts including temperature and relative humidity insurance to more than 11,000 farmers in over six Indian states.

Sources: World Bank (2003) and Bryla and Syroka (2007).

Page 99: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

99

Global climate insurance

The design and development of global climate insurance are fraught with difficulties, and efforts to launch climate insurance under the UNFCCC have so far been unsuccessful. Meanwhile, the World Bank is developing a system of bonds for governments to help ensure their financial liquidity in times of catastrophic loss. Currently, it may be better to proceed on an experimental basis outside the Convention rather than make attempts to negotiate an insurance package within the post-2012 agreements. It would nevertheless be helpful to develop these initiatives alongside the Convention negotiations and to prepare periodic assessments of progress.

Experience with natural hazard and disaster insurance provides some useful lessons and identifies some of the ways in which climate insurance might improve global welfare. There are some important differences, however, that suggest that insurance for climate-related weather events should be considered in a category of its own. Perhaps most important, the UNFCCC anticipates and proposes the use of insurance as a risk-sharing mechanism for dealing with the unequal impacts of a global environmental problem. In addition, the global atmosphere is a common property resource to be used for the benefit of humanity, so all can claim the right to a basic level of protection. Experience suggests that insurance can be a useful mechanism to achieve sound risk management and, if properly designed, international equity. Climate insurance is not likely to be made widely available, however, without some sort of public intervention. At the same time, the moral hazard involved in making insurance available at sufficiently attractive rates can be a significant problem. Taking all of these insights into account, therefore, it seems reasonable that any proposed scheme involving public intervention should include eligibility criteria.

Despite these complications, it is possible to imagine what a global climate insurance program (CIP) might look like. Its purpose would be to create a pool of financial resources that would compensate victims for climate-related losses. Losses could be reported at the national level, and they could be subject to international verification. Certain exceptions might be created, especially in cases in which specific weather criteria have been established (as, for example, in the case of weather derivatives). The financial pool could draw its initial funding from Annex 1 countries, and developing countries might subscribe to the program by agreeing to certain programmatic criteria. These eligibility criteria could be designed to ensure that the CIP could operate effectively within a subscribing country. The program could be established under an appropriate international authority, such as the UNFCCC, or one or more of the international financial institutions such as the World Bank or the International Monetary Fund.

What might the qualifying criteria be? Criteria at the national level might require specific climate risk management practices. These practices are likely to vary according to the nature of the climate hazard and the level of vulnerability. They might include building codes, land-use planning and regulation, forecasting and warning systems, and other climate adaptation measures that reduce vulnerability. They could even require that participating countries agree to reduce greenhouse gas emissions. National participants in the CIP could specify their preferred level of insurance and negotiate the types of climate risk coverage that they could achieve and their premiums.

Having subscribed to the CIP, national governments would adopt legislation and policies designed to make the program available to their citizens and institutions. Potential clients would include individual property owners, public and private enterprises, and possibly the national

Page 100: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

100

government itself. Marketing could be handled by local private insurance industries, and the policies could be written by an administering government agency or by the insurance companies under a “write-your-own” arrangement.

Such a CIP would offer many potential benefits. First of all, it would provide a mechanism for risk spreading and sharing open to the whole global community—an appropriate response mechanism for a global environmental issue like climate change. Second, it could promote, and to some extent require, clients and beneficiaries to implement adaptation measures. It would thereby contribute in a verifiable way to the long-term reduction of vulnerability, and it would counteract social forces that currently encourage maladaptation and otherwise enlarge overall vulnerability through unsustainable development practices. By reducing vulnerability, the CIP would simultaneously slow and perhaps eventually reverse the growing costs of emergency disaster relief and humanitarian assistance. If successful, the CIP would slow the growth in disaster and other climate-related losses, improve the balance of payments of developing countries, and thus diminish their need for debt-financed reconstruction. Reduced borrowing would mean reduced debt burdens and reduced likelihoods of destabilizing debt crises. At the same time, the CIP would remove some of the most important obstacles to achieving the Millennium Development Goals by consolidating development aid for the planet’s poorest people and poorest countries. Finally, it would provide a mechanism by which developing countries would benefit from committing themselves to emissions targets and timetables.

5.3 Strengthening important ongoing development initiatives to support climate change adaptation and mitigation

5.3.1 Secure property rights

Meeting the challenges of climate change adaptation in agriculture requires long-term investment by farmers. Long-term investments (such as integrated soil fertility management, tree planting, and water harvesting) require secure property rights to provide people with the incentive and authority to make the investments (Meinzen-Dick et al. 2002). By changing the profitability of land, such as through the potential for income from carbon markets and biofuels, climate change may also worsen the position of famers with insecure property rights, leading to expulsion from their land as landlords seek to increase their share of the new income streams. Improvement in land rights is therefore an essential component in effective and equitable adaptation.

Secure property rights do not necessarily have to be individual or titled land; secure collective or customary tenure can also be sufficient (Bruce and Migot-Adholla 1996; Sjaastad and Cousins 2008). In cases where pressure on land is growing, however, customary tenure may no longer be secure. These cases call for innovative approaches to securing land tenure, which may involve alternatives to titling. These alternatives could range from recognizing customary rights to land, to identifying agents to represent customary interests, to formalizing groups and granting them collective rights over resources (Fitzpatrick 2005; Kanji et al. 2005).

Climate change is making water access inherently less secure because water flows are becoming less predictable. The declining availability and increasing variability in rainfall and streamflow in many regions will decrease the security of water access. Therefore, it is increasingly important to influence other factors that reduce security of access, especially lack of secure water rights. Secure water rights empower users by requiring their consent to any

Page 101: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

101

reallocation of water and granting users compensation for transferred water. Secure and well-defined water rights give users incentives to invest in water-saving technology. A system of tradable water rights can also encourage users to consider the full opportunity cost of water, including its value in alternative uses, thus providing incentives to economize water use and gain additional income by selling saved water. Moreover, a properly managed system of tradable water rights will give water users incentives to internalize the social and environmental costs imposed by their water use, reducing the pressure to degrade resources.

5.3.2 Agricultural policies

Whereas a supportive policy environment is a critical element of success, policy-induced distortions can lead to major problems. Improving economic incentives for adaptation thus requires, for example, removing the existing perverse subsidies on water, energy, and fertilizer that encourage environmentally damaging overuse of these resources and inputs. A key lesson from the Green Revolution is that its negative environmental outcomes were largely due to a poor policy environment characterized by inappropriate subsidies and biased pricing and trade policies that favored monoculture rice at the expense of more diversified farming systems (Pingali and Rosegrant 2001).

Climate change should be mainstreamed in all agricultural policies to limit policies and investments that inadvertently encourage, rather than minimize, vulnerability to the impacts of climate change. Mainstreaming climate change in agricultural policies would help avoid investments in agricultural research and development (R&D) for crops that are not likely to thrive under global warming in certain Asia-Pacific environments, investments in agricultural water management technologies that perform poorly with increased temperature (such as sprinkler as compared with drip), and investments in livestock expansion in areas expected to experience declines in pastures and grazing lands as a result of climate change, such as Inner Mongolia and Mongolia. Similarly, mainstreaming climate change will help focus agricultural policies toward enhanced resilience under extreme events and global warming.

5.3.3 Trade policies

Trade liberalization is an important adaptation strategy because producing food based on local comparative advantages regarding resource availability will help reduce GHG emissions and allow countries to adapt to climate change more effectively and efficiently. Growing scarcity of water, fuel, and land has the potential to drive up food prices, limiting access to food. The experience in 2007–08, when several countries imposed trade restrictions as a result of higher food prices and increased price volatility, shows how breakdowns in trading systems can raise potential threats and adverse impacts for food security. Thus, restoring confidence in international trading systems will be crucial (Box 5.6). Effective food trading systems will also require continued advancements in food safety standards, both through the Codex Alimentarius and through enhanced risk analysis and risk management.

But will smallholders in the Asia and Pacific region be able to benefit from increased trade liberalization? Cooperative storage and contract farming both for export and for local supermarkets with growing retail shares in developing Asia are important means of increasing certainty and stability in smallholder agricultural production.

Page 102: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

102

Box 5.6. Restoring confidence in international agricultural trade

The ongoing failure of the World Trade Organization’s (WTO) Doha Round, together with the sharp increase in food prices that stimulated export bans and other restrictions of trade by many countries, has resulted in declining confidence in agricultural trade. The restrictive agricultural trade policies adopted by several developing countries also undermine the benefits of global integration, adding to the distortions already created by rich countries’ longstanding trade policies. Agricultural globalization is regressing, with adverse effects for the poorest countries. Rule-based, fair, and free international trade is particularly critical in times of crisis, as the export ban problems underline. A sound global trade system is especially crucial in the context of climate change. As shown in Chapter 2, the impacts of climate change on agricultural growth and production will make developing countries increasingly reliant on food imports. To increase confidence in international agricultural trade, the WTO Doha Round should be completed; OECD countries should reduce or eliminate trade restrictions that limit developing-country export access to markets; and buffering mechanisms should be established to better address volatility in world markets. Alternative or complementary approaches to market stabilization for cereals include a joint pooling of fixed portions of national stocks into an international grain reserve, and/or a financial facility, provided by the International Monetary Fund (IMF), for imports by countries in food emergencies.

5.3.4 Other environmental policies

In addition to secure property rights, farmers and land and water managers need not only incentives to make the decisions to sustain these resources, but also sufficient flexibility to adapt efficiently to climate change signals. Market solutions that promote sustainable natural resource management and mitigate the negative impacts of climate change are a potential method to reduce emissions and improve soil fertility and productivity and water efficiency while improving livelihoods of poor communities in developing countries.

With rising food, energy, and land prices—and in the longer run carbon prices—it is necessary to overcome past constraints and fully implement green markets, such as the Clean Development Mechanism (CDM). Under climate change, rising energy prices will change the relative effectiveness of different types of irrigation and water allocation policies. Higher energy prices will increase the cost of poor distribution systems and place restrictions on water trading and water subsidy regimes (Zilberman et al 2008). The increasing cost of water and energy subsidies worldwide will lead to significant pressure and increased incentives to reform water management to improve water use efficiency, including using water markets or other economic incentives, reducing subsidies, and making targeted investments in efficiency-enhancing technology.

With rising input and output prices, efficiency pricing of water and markets in tradable water rights is an important component of strengthening climate change adaptation, because it improves water use efficiency across sectors. Large-scale adoption of water markets or efficiency pricing of water is challenging, however, and will require innovative designs to protect farm incomes—for example, brokered trading to ensure fair compensation for irrigators who trade water. Appropriate pricing systems in the domestic and industrial sectors can enhance efficiency and equity of use, target subsidies to the poor, cover delivery costs, and generate adequate revenues to finance the needed growth in supply coverage. Pricing policies for the irrigation sector are inherently more difficult to realize because of political concerns, complex design and implementation, and potentially adverse impacts on poor consumers and farmers.

Page 103: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

103

Excessively high water prices are likely to severely reduce farm income. Moreover, in much of the developing world, irrigation consists of large systems serving many small farmers. Measuring and monitoring deliveries to this large numbers of end usersas would be required for volumetric chargesis too costly. Despite these difficulties, water-pricing systems, such as a water brokerage system, can be designed to introduce incentives for efficient water use, recover operations and maintenance (O&M) costs, and at the same time protect or even increase farm incomes. In a brokerage system, a base water right is established at major turnouts to individuals, groups of water users, or water user associations, which regulate distribution within the group. A fixed base charge would be applied to the initial (historical) quantity, sufficient to cover O&M and longer-term asset replacement (depreciation) costs. The brokerage agency—for example, a river basin authority—would then broker water trades. For demand above the base water right, an efficiency price would be agreed upon equal to the value of water in alternative uses; for demand below the base right, users would be compensated at the same price for unused water. Reform of water-pricing policy in developing countries faces many technical, administrative, and political constraints, but with increasing water scarcity under climate change and declining financial resources available for irrigation and water resource development, such reform is essential (Rosegrant and Cline 2002).

Existing markets favor production of crops or livestock relative to production of environmental services. Payments for environmental services (PESs) can help reflect the value of environmental services more accurately and thus enhance their production. Payments compensate farmers for the costs they bear in producing these services (FAO SOFA 2007), giving them incentives to invest in land use practices that can increase and diversify their income streams and help them both adapt to and mitigate climate change. It is an important option to consider for several other reasons as well.

First, farmers are the largest group of ecosystem managers on earth, and they have an important role to play in improving the management of global and local natural resources. Second, paying farmers for environmental services can be a relatively inexpensive and quick means of responding to some environmental problems. Third, environmental service payments can be a more equitable way to manage environmental problems, particularly when poverty is a cause of environmental degradation. Payments for environmental services provide one option to offset pressures to generate biofuel benefits out of agricultural ecosystems at the expense of environmental services. Policies and contract reforms should be implemented to bring smallholder farmers—who have often been bypassed because of property rights issues and high transaction costs—into PES systems (FAO SOFA 2007).

PES approaches may be most effective when local communities are involved in negotiations to determine the terms of the payments. For example, downstream users in a watershed may try to negotiate with upstream users to protect the water from pollution and sedimentation. The downstream users may offer a payment or reward in exchange for implementing agreed-upon management practices. When the initiative comes from local people who are direct stakeholders, it may make sustainability easier to achieve, because the downstream users will have an interest in continuing to monitor compliance (Pender 2006). Such negotiation and collective agreements are more likely within relatively small and cohesive communities than they are between communities, and the ability to ensure that all resource users benefit is greater. The village of Sukhomajri in India is one of the best examples in which the benefits of a locally initiated watershed development effort were broadly shared in the

Page 104: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

104

community in return for compliance with grazing restrictions, leading to dramatic improvements in natural resource management, household food production, and livelihoods (Dixon et al. 2001). Lack of replication of this experience may be less dependent on local leadership and other idiosyncratic factors and more related to absence of local control over resources.

5.3.5 Social protection

Given the low levels of income and savings in poorer communities, as well as the weak economic position of certain states, developing countries will need to develop more robust social protection schemes at the individual and national levels. At the individual level, such measures can include employment programs, cash transfers, and weather- and crop-related insurance. At the national and regional levels, countries will need to further leverage international financial markets and develop relationships with the financial services sector to pool and transfer their risk to ensure that they will not have to significantly redirect national budgets in cases of climate shock.

Comprehensive social protection initiatives are required to address the risks facing the poor owing to climate change and increasing climate variability. Appropriate social protection interventions include both protective measures to mitigate short-term risks and preventative measures to preclude long-term negative consequences. By protecting against downside risk, effective social protection also reduces risk aversion in farmers’ production decisions, enhancing the potential for adaptive farming systems. Introducing or scaling up these interventions is, however, complex, expensive, and dependent on a country’s knowledge base and capacity (IFPRI 2008).

At the core of the protective measures are conditional cash transfer programs, pension systems, and employment programs. These programs exist in many low-income countries and should be scaled up. Where such interventions do not exist, countries should introduce targeted cash transfer programs in the short term. If food markets function poorly or are absent, however, providing food is a better option. Microfinance, which includes both credit and savings, will allow the poor to avoid drastic actions such as distress sales of productive assets that can permanently damage future earning potential. Furthermore, Francisco (2008) has suggested the potential for developing microinsurance index-based schemes in Southeast Asia. Partnerships among international organizations, national governments, nongovernmental organizations, and the private sector should examine schemes that have worked well in some countries and pilot test them in the region.

Preventative health and nutrition programs targeted to vulnerable population groups (such as mothers, young children, and people living with HIV/AIDS) should be strengthened and scaled up to ensure universal coverage. This measure is essential to prevent the long-term consequences of malnutrition on lifelong health and economic productivity. In addition, school feeding programs can play an important role in increasing school enrollment, retaining children in school, and enhancing their academic achievement.

Overall, expected results of social protection programs include preventing long-term adverse consequences of early childhood malnutrition, increasing protection of assets, and maintaining school participation rates. Many of these actions should take place at the national level, but many countries lack the resources to implement them. Donors should expand support for such programs in conjunction with sound public expenditure reviews (IFPRI 2008).

Page 105: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

105

5.3.6 Financial markets: The role of microfinance

Microfinance services (MFSs) can be an important tool in reducing the vulnerability of the poor and, in the context of climate change adaptation, can provide poor people with the means to diversify, accumulate, and manage the assets needed to become less susceptible to shocks and stresses or to better deal with their impacts. Yet these benefits may not apply to everybody. MFSs typically do not reach the chronically poor, may encourage short-term coping at the expense of longer-term vulnerability reduction, or even increase vulnerability. These limitations and risks aside, MFSs can still play an important role in vulnerability reduction and climate change adaptation among some of the poor, provided services better match client needs and livelihoods (Hammill, Matthew, and McCarter 2008).

Hammill, Matthew, and McCarter (2008) note that microfinance services can be divided into three main types. Microcredit lends funds to poor people so they can exploit their capacities for income production (job creation, enterprise growth, and increased production) and is about asset building and diversification. Returns are consumed, saved, and/or reinvested. Loans are also offered for nonproductive purposes that may contribute to reducing vulnerability, such as emergency loans, education loans, and home improvement loans. Microinsurance (Pierro and Desai 2008) protects poor people against specific perils (such as injury, death, and natural hazards) in exchange for regular premium payments (Churchill 2006). Thus, like the social protection policies already described, it protects assets and gives people the freedom to pursue profit without fear, ideally leading to increased income production and adaptability (Morduch 2006). Microsavings are small balance deposits for the safe storage of money that allow people to obtain lump sums to meet both predictable and unpredictable expenses. They can be used as insurance or for investment, yielding the same results for asset bases already described (Hammill, Matthew, and McCarter 2008).

Potential pitfalls need to be avoided. If microfinance is essentially a coping mechanism, it is not likely to be a pathway toward adaptation and could even increase vulnerability. Debt burdens can also increase to unsustainable levels. Furthermore, if governments see microfinance as a substitute for appropriate levels of social protection, the adaptive effects could weaken. If these pitfalls can be avoided, the most powerful case for MFSs with regard to climate change adaptation is their ability to help families build and diversify assets so that they have more than one means of livelihood or more than one skill set to avoid dependency. Green microfinance, through service conditions that provide sustainable resource stewardship, may reinforce longer-term vulnerability reduction gains. For example, a partnership between the Self-Employed Women’s Association (SEWA) Bank and SELCO-India (a social enterprise providing sustainable energy solutions and services) seeks to meet the energy needs of self-employed individuals and microenterprises for processing, agriculture, and other livelihoods (McKee 2008). Although the need for green microfinance is recognized, appropriate terms and modalities need to be developed to make it effective without sacrificing positive social impacts. Balancing quick gains and short-term loan repayment schedules with longer-term sustainable management practices will continue to challenge the industry (Hammill, Matthew, and McCarter 2008).

5.3.7 Disaster preparedness

Coastal defense systems will be crucial for disaster preparedness in Bangladesh, Viet Nam, and many of the island states in the Pacific. These investments require attention to financial and

Page 106: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

106

human resource capabilities at both the development and the maintenance stage. For examples the coastal sea dike defense system in northern Viet Nam degraded significantly following decollectivization, as the management authority for maintenance was shifted from agricultural cooperatives to decentralized communes, which considered aquaculture development a higher priority (Adger 2001). In the Viet Namese Mekong Delta, mangrove replanting has been a key component of the coastal defense system since the late 1990s. The mangroves not only provide physical protection and environmental sustainability, but also generate ecosystem goods and services (Tri et al. 1998; Adger 1995). According to studies by CARE (2007), a 100 m-wide band of mangrove forest in coastal Viet Nam was sufficient to reduce the amplitude of tidal waves by 50 percent and the energy by up to 90 percent. A comparison of a 1996 and 2005 typhoon found that there were significant improvements, no loss of human life, and a significant drop in property damage, and mangrove survival was 63 percent.

Funds have been made available through the Indian Ocean Tsunami Warning and Mitigation System (IOTWS) through the Economic and Social Commission for Asia and the Pacific (ESCAP). The aim of the fund is to build tsunami early warning capabilities through the building of institutional, technical, and systemwide capacity (ESCAP 2009). The fund will be administered by governments to allow for their own design and implementation of projects and priority identification. At the end of 2008, the fund had approved 11 projects in the region with a budget of US$9.2 million.

5.4 Implementing Climate Change Adaptation Policies

5.4.1 Mainstreaming climate change and adaptation into development planning

Development policies and plans at all levels need to consider the impacts of climate change on the agricultural sector. National and regional policymakers must integrate the effects of climate change and the outcome from assessments and scenarios into their national plans and policies in the agriculture sector. Advanced planning, or “climate-proofing,” will ensure that climate change does not disrupt or render ineffective development plans that are critical to at-risk or vulnerable communities with low levels of development. Moreover, mainstreaming should aim to limit development policies and plans that inadvertently encourage, rather than minimize, vulnerability to the impacts of climate change. Many of the aforementioned adaptation strategies are already part of sound development policy advice, which should make mainstreaming easier. At the same time, adaptation to climate change should be recognized as a critical element of development policy that will require both innovative new ideas and additional funding commitments, and this reality should not be lost in mainstreaming efforts. Funding options are described in greater detail later.

Although the interdependence of climate change adaptation and sustainable development should be self-evident, it has been difficult to combine them in practice. A significant adaptation deficit exists in many developing countries, particularly those populated by the rural poor who rely on agriculture for their very subsistence. Although the UNFCCC includes clearly defined objectives, measures, costs, and instruments for mitigation, it does not do so for adaptation. Agrawal (2005) reports that much less attention has been paid to how development could be made more resilient to climate change impacts and identifies a number of barriers to mainstreaming climate change adaptation within development activities. These barriers include segmentation and lack of coordination within governments and donor agencies, the lack of

Page 107: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

107

relevant climate information for development-related decisions, and perceived trade-offs between climate and development activities.

Despite these barriers, the development community recognizes the linkages between development and climate change adaptation. Schipper and Pelling (2006) note that climate change has been identified as a serious risk to poverty reduction in developing countries, particularly because these countries have a limited capacity to cope with current climate variability and extremes, not to mention future climate change. Adaptation measures will need to be integrated into poverty reduction strategies to ensure sustainable development, and this goal will require improving governance, mainstreaming climate change measures, and integrating information on climate change impacts into national economic projections. Based on case studies of natural resources management in Bangladesh, Fiji, and Nepal, Agrawal (2005) recommends several priority actions for overcoming barriers to mainstreaming, such as screening projects for climate-related risk, including climate impacts in environmental impact assessments, and shifting the emphasis from creating new plans to better implementation of existing development measures and policies.

To mainstream climate change adaptation, countries will need to undertake multifaceted risk assessments that incorporate not only climate risk, but also existing vulnerabilities such as low levels of development, poor governance, and political instability and expected future trends such as population growth, rapid urbanization, and increasing water scarcity. Qualitative and quantitative scenarios will need to be developed at the country level and potentially at the subnational and regional levels. Combined with detailed economic analysis of adaptation options, these multifaceted risk assessments and scenarios should serve as the basis for developing comprehensive and robust adaptation plans. The National Adaptation Plans of Action (NAPA), with the financial support of the UNFCCC acting through the Global Environment Facility (GEF), could be key mechanisms for mainstreaming climate change into development planning, but progress on NAPA has been slow.

5.4.2 National Adaptation Plans of Action (NAPA)

All countries, as part of their responsibilities under the UNFCCC, should create national adaptation plans. These plans would take a broad strategic view of the future development path of the country and consider how it could best be designed or modified in light of expected climate change. Within such a strategic view, policies for sectors and regions could be examined and adjusted to account for climate change. Sectoral policies would likely include those for agriculture, forests and fisheries, water and other natural resources, health, infrastructure, and ecosystems. In addition, the policy review could include the management of extreme events such as droughts, storms, and floods and areas of particular risk such as exposed coastal zones and steep mountain slopes. Specific adaptation measures could then be evaluated and selected within the context of a climate-sensitive strategy and set of policies.

Financing for these plans, however, is limited to the least-developed countries, and NAPAs are not comprehensive adaptation plans but are confined to urgent or priority measures (see Table 5.4 for costs of financing NAPAs). A common concern of the developing countries has been that their participation in multilateral environmental agreements imposes costs on them as they undertake new obligations to address global environmental problems created to a large extent by the industrialized countries. It seems realistic therefore to suggest that the developed countries, acting collectively through the GEF, should support the preparation of NAPAs. This

Page 108: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

108

step would not only help to ensure that climate is adequately considered in national development plans and sectoral policies, but also to reassure donors and investors that climate change adaptation measures are well conceived and represent sound expenditures.

The preparation of NAPAs is only an early step in moving countries toward an effective adaptation response. The plans need to be implemented, and further support for this implementation will clearly be required. Most of the present funding for adaptation has been on a voluntary basis. The GEF has established funds or “windows” to which developed countries make contributions and from which developing countries can obtain funds indirectly through one of three implementing agencies (the UN Development Programme, the UN Environment Programme, or the World Bank). The growth in these funds has been slow, partly because the donor countries seem to lack sufficient confidence that the modalities for the effective use of the funds exist. Negotiating the details of the preparation of the plans will be time consuming, and thus action must be taken if such ideas are to be included in post-2012 agreements. It may be appropriate to proceed slowly in developing NAPAs, however. If NAPAs are truly to be comprehensive and part of the mainstreaming process, it may be more effective to build them into Poverty Reduction Strategy Plans or other national development plans currently coordinated by multiple donors. Without this integration, they may simply add another layer of planning rather than aiding the mainstreaming process.

5.4.3 Significant new investments

Much of adaptation in the agricultural sector can be implemented without huge new investments, but some key initiatives, such as agricultural research, will require significant new investments. Increased and diversified investments are needed in plant breeding, livestock improvement, and other interventions at the biological and molecular levels to enhance agricultural productivity in ways that ultimately contribute to poverty reduction, agricultural development, and economywide growth throughout the region. Such a program requires heavy investment in advanced scientific expertise and equipment, as well as a political and social commitment to long-term funding of agricultural science and technology at levels significantly greater than current funding. Furthermore, it requires new investments to create organizations that are more dynamic, responsive, and competitive than the public organizations that make up the bulk of national agricultural research and extension systems in Asia today.

Major investments in water infrastructure are also needed. Dams have proven to be an effective way to protect agricultural systems and human settlements from water variability, and a higher demand for dams is expected as a result of increasing water variability and energy demand. Big dams are known, however, to cause considerable environmental and social impacts. Furthermore, investment is needed in engineering techniques to reduce environmental impacts, management techniques to optimize dam use, planning tools to reduce social impacts, and tools to improve design and operational techniques. Investments should also be made to scale up underground storage techniques. Finally, more investments should be made in research to assess the viability of inter-basin transfer schemes, which can be politically challenging and risky in light of future uncertainty about water availability.

Human-made or natural infrastructure to protect against sea-level rise is another important area for new investments to adapt to climate change. As for all new investments, it is important to ensure that resources, both financial and human, are sufficient not only to develop the structures, but also to maintain them in the long term.

Page 109: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

109

5.4.4 Cost of adaptation for the Asia-Pacific region

Global adaptation costs

Despite the proliferation of adaptation funding windows, most of the activities funded are for mitigation, not adaptation. For example, the GEF climate change operational program funded activities worth nearly US$1 billion, but only a small number were directed toward adaptation (Huq and Burton 2003). TERI (2006) discusses some difficulties in implementing adaptation activities:

it is difficult to obtain baseline information for incremental calculation of costs;

funding agencies require presentation of “global environmental benefits,” but such benefits from adaptation projects can be expressed only at local and sometimes regional levels;

adaptation to future climate change must be separated from activities that enhance adaptation to climate variability; and

most often, adaptation activities are closely linked to other aspects of development, making it difficult to determine the adaptation component of a project (IISD 2004).

The inadequacy of these efforts is shown clearly in comparison with the estimated costs of adaptation worldwide. At the global level, several estimates of the required financing have been undertaken, most recently by the UNFCCC and in the Human Development Report 2007/2008.

The UNFCCC estimated the annual investment flows needed on a sectoral level for adaptation in 2030, as follows:

Agriculture, forestry, and fisheries: US$14 billion

Water resources: US$11 billion

Human health: US$5 billion

Coastal zones: US$11 billion

Infrastructure: US$8–130 billion

According to these estimates, global adaptation efforts will require US$49 to US$171 billion a year in 2030.

The Human Development Report 2007/2008 used a different methodology, focusing on three main categories of financing requirements. The HDR estimated that the following new financial flows will be required on an annual basis in 2015:

Climate-proofing development investment: US$44 billion

Adapting poverty reduction to climate change: US$40 billion

Strengthening disaster response: US$2 billion

In total, the HDR estimates that adaptation will necessitate additional financial flows of US$86 billion a year in 2015.

Page 110: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

110

To put both of these estimates in perspective, the OECD calculated that a total of US$103.7 billion was spent on official overseas development assistance (ODA) in 2007.16 The HDR estimate of US$86 billion and the rough average of the UNFCCC estimated range, US$110 billion, are within range of this figure; however, this would require that all ODA be used for climate change adaptation. Irrespective of the accuracy of these figures, adaptation will clearly require moving beyond the traditional development aid paradigm and necessitate the development of new and innovative financing solutions. In addition, adaptation needs and poverty reduction goals will need to be integrated into broader economic development to make the best use of scarce funds.

Alongside the ongoing work under the auspices of the UNFCCC, most actors in the international development and humanitarian community, as well as select private sector firms, have begun their own adaptation efforts. These efforts are of two distinct types—(1) mainstreaming climate change impacts into existing program portfolios and (2) developing new and additional activities in the area of adaptation.

Adaptation costs for the Asia-Pacific region

For the Asia-Pacific region, adaptation costs are expected to be particularly high to combat sea-level rise. Protecting South and Southeast Asia from sea-level rise (greater than 50 centimeters) could cost US$305 billion alone (Tol 2002). Table 5.5 presents estimates of costs as a percentage of GNP per year for global regions. Small island nations in the Pacific and Indian Oceans will face the largest burden (approximately 0.75 percent of GNP per year), followed by coastal communities along the Indian coast of Asia (0.52 percent of GNP per year) (Francisco 2008). Over 27 million people who live along the Indian coast would be at risk, which are estimated to reduce to 3 million people with these estimated adaptation investments (Francisco 2008).

Table 5.4 shows the costs of priority activities of adaptation in selected LDCs in the Asia and Pacific region. The projects in the five countries listed—Bangladesh, Bhutan, Cambodia, Samoa, and Tuvalu—are estimated to cost US$72 million, which would be double the regional funds currently available in the Climate Change Fund or CCF. Rehabilitating the upper Mekong to reduce flooding risks is the most expensive measure—at least 5 times the average costs of other projects—totally US$30 million. Maintaining water resources is a priority activity for most of the countries listed, including flood protection, rainwater harvesting, and irrigation.

5.4.5 Financing adaptation

In recent years, new mechanisms have been established to support adaptation, including the Least Developed Country Fund, the Special Climate Change Fund, and the Adaptation Fund. Huq (2002), Brander (2003), Desanker (2004), Huq (2006), and Huq, Reid and Mussay (2006) trace their evolution. They have provided the opportunity to mainstream adaptation into local and regional development activities. One critical problem with mainstreaming climate change adaptation with existing development assistance has surfaced, however. The boundary between existing development assistance and the additional funds promised under the UNFCCC for adaptation is vague. This ambiguity may require difficult decisions about how much of an

16 This figure represents official development assistance from Development Assistance Committee (DAC) countries, which is

meant to account for over 80 percent of all development assistance. This figure does not include contributions from non-DAC OECD countries or grants by private voluntary agencies and foundations in DAC member countries.

Page 111: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

111

adaptation project is for “regular” development and how much is specifically for adaptation to climate change.

This distinction is important because it carries implications about the distribution or allocation of costs for particular actions within UNFCCC mechanisms such as the GEF. For example, Burton (2004) and Huq and Reid (2004) note that calculating the costs of adapting to future climate change (as opposed to current climate variability), as well as the local nature of resulting benefits, are both problematic vis-à-vis the GEF requirement for calibrating global environmental benefits. On the other hand, opportunities also exist. Under the UNFCCC and its Kyoto Protocol, three funds have been created to support adaptation to climate change. The GEF operates the Special Climate Change Fund (SCCF) and the Least Developed Country Fund (LDCF). Institutional arrangements to operate the Adaptation Fund under the Kyoto Protocol are under discussion and are expected to be finalized by COP15 in Copenhagen, December 2009 (the 15th meeting of the Conference of the Parties [COP] of the UNFCCC).

COP guidance on GEF support for adaptation identifies three stages. Stage I provides support for the national communications process, a portion of which is the vulnerability and adaptation assessment. Stage II provides further assistance for other capacity-building efforts for adaptation. Stage III refers to support for actual adaptation activities, including insurance, and has been implemented in the form of the GEF Strategic Priority on Adaptation (SPA). The GEF has allocated US$50 million under SPA, of which US$5 million is devoted to piloting community adaptation initiatives through the Small Grants Program (SGP).

The goal of the Community-Based Adaptation (CBA) Program is to pilot the community component of the GEF and to provide the basis upon which the GEF and other stakeholders can effectively support small-scale adaptation activities. This goal will be realized through three immediate objectives: (1) development of a framework—including new knowledge and capacity, spanning the local to the intergovernmental levels (cross-scale “policy laboratories”)—to respond to unique community-based adaptation needs; (2) identification and financing of diverse community-based adaptation projects (small-scale “policy laboratories”) in a number of selected countries; and (3) capture and dissemination of lessons learned at the community level to all stakeholders, including governments.

The LDCF was established to support the preparation and implementation of NAPAs. The operational modalities and procedures have been finalized, and one project for Bhutan has already been approved under this fund. At present the LDCF has approximately US$115 million for funding priority activities in 48 least-developed countries (LDCs) under the UNFCCC. Bangladesh, Bhutan, Cambodia, Kiribati, Samoa, Tuvalu, and Vanuatu are 7 of the 48 official LDCs that have prepared NAPAs and are therefore currently eligible for funds (GEF 2008). Cambodia, however, is the only country receiving funds—approximately US$1.9 million from the LDCF specifically in the agricultural sector (GEF 2008). The SCCF was established to finance developing-country activities in adaptation, technology transfer, key sectors (energy, transport, industry, agriculture, forestry, and waste management), and economic diversification for countries dependent on the fossil fuel sector. The Adaptation Fund is intended to fund concrete adaptation projects and programs in developing countries that are particularly vulnerable to the adverse effects of climate change. The funding is provided by a 2 percent levy on proceeds from Clean Development Mechanism (CDM) projects (excluding those undertaken in LDCs) and “other sources.” The total scale of the Adaptation Fund will therefore depend on the volume of activity in the CDM.

Page 112: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

112

Although there has been a great deal of attention to this issue over the past year, much of the related activity by international actors has focused on the first type of adaptation action—mainstreaming climate change into existing program portfolios. According to the OECD, however, there has been little concrete progress even in this area. Although awareness of climate change impacts has increased significantly and several tools have been developed to support “climate proofing,” few development programs have integrated the impacts of climate change into their plans.

Beyond the efforts being undertaken to “climate proof” existing portfolios, most international development and humanitarian agencies have only recently or are in the midst of developing their own individual strategies for additional new activities in the area of adaptation. Though initiatives in these areas have recently proliferated, few concrete activities are currently underway, especially at the country level or below. Examples of some noteworthy exceptions include impact analysis by research institutes such as the Columbia University Earth Institute and The Energy and Resources Institute (TERI); country-level support by UNDP, UNEP, and other UN line agencies; the economic case for adaptation and engagement with the insurance industry being developed the World Bank; and the US$70 million Climate Change Initiative financed by the Rockefeller Foundation.

Specifically in the Asia and Pacific region, the ADB is supporting the creation of regional funding modalities. The main mechanism in the region available for both adaptation and mitigation is the Climate Change Fund (CCF), with an initial contribution of US$40 million (Sharan 2008). Two other smaller funding sources have been created, the Water Financing Partnership (WFPF) and the Poverty and Environment Fund (PEF). The WFPF has secured donor commitments totally US$26 million, while the PEF has a more modest US$3.6 million budget (Sharan 2008).

In addition, the private sector—the insurance and reinsurance industries in particular—has started to engage in adaptation activities in developing countries. The most advanced initiatives have been developed by two global reinsurance companies, Munich Re and Swiss Re. These initiatives focus on developing new risk-transfer products such as microinsurance, weather and crop insurance, and other mechanisms such as risk pooling and disaster-related bonds. A set of pilot programs is currently underway in various developing countries, and implementing partners are assessing their efficacy and the overall business case for engagement.

5.5 Reforming climate-change related governance and institutions

Effectively planning and implementing climate change adaptation for agriculture will be based on a number of factors. First, it will require the engagement of a core ministry, such as the Ministry of Finance or Planning, alongside the Ministry of Agriculture, to ensure strong government support (Stern 2006). Such engagement has been developed in many cases for climate change mitigation, but it has not often extended to adaptation. Second, the core capacities of developing-country governments will need to be further developed. Such capacity building is required across a number of areas, including technical subjects such as climate forecasting and scenario planning, as well as general development topics such as governance, accountability, and empowerment of local communities.

Third, adaptive and flexible management will be essential, including the capacity to monitor the results of managers’ decisions and to modify actions as needed. The broadening

Page 113: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

113

nature and increasing severity of potential climate impacts in a given area and the unavoidable uncertainties associated with predicting these impacts requires innovative approaches to management and development that go beyond centralized prediction and control practices (Nelson et al. 2008; Pahl-Wostl 2007a). One approach—adaptive management, or adaptive governance—has received attention because it enables decision makers and resource managers to work with the inherent uncertainty associated with climate change (Pahl-Wostl 2007b; Brunner et al. 2005; Tompkins and Adger 2004; Folke et al. 2002).

Although interpretations of adaptive management and governing institutions often differ between disciplines (Stankey et al. 2005), such institutions have several defining characteristics. First, the management scale is often realigned with the scale of ecological processes (for example, the watershed or the ecosystem) (Cumming et al. 2006). Second, they are based on a local or regional community-based management system (Olsson et al. 2004). Third, they involve collaboration and integration of various organizations and institutional arrangements at all scales of decision making to foster flexibility, balance divergent interests, and promote coordination and deliberation among diverse stakeholders (Folke et al. 2005; Dietz et al. 2003). And fourth, an adaptive governance approach requires that managers be knowledgeable about scientific and local information, as well as the implementation of policy experiments that develop understanding, prioritize learning as an objective, and improve the ability to manage uncertainty (Lee 1999; Holling 1978). This experimental approach goes beyond trial and error, for it takes an explicitly scientific approach to testing and subsequently learning from empirically informed management decisions (Arvai et al. 2006).

As stated at the beginning of this chapter, both mitigation and adaptation response options need to be implemented by a variety of actors at the local, community, national, and global levels. To decide which actor is most appropriate, it is useful to look at the time and spatial dimensions involved in the specific adaptation (and mitigation) response. If solutions in the long run are to be sustainable, involving local people, considering context-specific issues in local policies, and recognizing the increasing role of international institutions in multicountry agreements are essential. At all levels, scaling up adaptation or mitigation policies requires the involvement of the private sector, because available funds are primarily in private agencies, and it is important to build on their successful strategies. Coordination among institutions becomes increasingly important, especially with high demand for better cross-sectoral planning tools and flexible and adaptive management systems.

Markets also play a coordination function, ranging from the global to the local. The question of when market (rather than state or collective action) institutions work best depends not so much on scale but on issues of transaction costs and attitudes toward markets. Working with many small suppliers of carbon “services” entails higher transaction costs than working with a few large-scale suppliers, which means that markets tend to favor plantations, for example, over smallholder agriculture or forest communities. Asymmetrical information, either about the actions of farmers or the benefit streams they could tap, will also hinder market-based coordination. Finally, the acceptability of market approaches will depend on values and attitudes toward resources and markets. For example, certain indigenous peoples’ groups have objected to the commoditization of their land and its carbon, which they feel has heritage value, whereas others may see markets as an opportunity.

In practice, many issues require policies and action at all of these levels and sectors. For example, effective REDD (Reduced Emissions from Deforestation and Degradation) agreements

Page 114: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

114

will require international market mechanisms to match those who wish to pay to offset their emissions with those who will sequester carbon, national governments that will broker agreements (for example, through a Designated National Authority [DNA] as currently employed for CDM agreements), and collective action groups to monitor compliance among local smallholders. Although local collective action can provide an effective means to measure and ensure compliance, whether the groups will have the motivation to execute this role on an ongoing basis will depend on whether the incentives exist. Continued participation is more likely if the group has been involved in the negotiations, has had a say in setting the rules, and receives a substantial benefit, either for the group or for its members.

Experience with collective action in other types of natural resource management suggests that systems that are developed in a top-down manner and do not engage local people in designing rules and systems are unlikely to create viable institutions that operate at the local level in the long run. This experience serves as a warning against focusing only on national-level negotiations and systems for climate change mitigation or adaptation, because they are unlikely to create effective institutions to execute the programs, especially among smallholders (Meinzen-Dick et al. forthcoming). A range of national and local institutions, public and private, are therefore needed. Rather than focusing exclusively on any single type of institution, policies need to develop polycentric governance arrangements within which multiple institutions play a role (Ostrom 1999). This situation also calls for coordination among different institutions.

5.5.1 Civil society

Farmers and villagers are likely to be affected by adverse climatic changes, and thus they may voluntarily develop and apply adaptation measures together by contributing their time and resources (Francisco 2008). This kind of risk-sharing practice constitutes community-based adaptation activity, one example of which is the adaptation project implemented in the Thua Thien Hue province of Viet Nam. Box 5.4 describes this successful project. Because communities worked together from the planning to implementation stages, the adaptation strategies fit their needs well. The project also shows that empowering civil society to participate in the assessment process, including identifying adaptation strategies and implementing the activities based on the plan, reduces the vulnerability of communities to climate change (Francisco 2008; Sperling 2003). Similar types of adaptation schemes should be tested in other Asian and Pacific countries subject to annual flooding.

Furthermore, in Cox Bazar, in eastern Bangladesh, when women became fully involved in disaster preparedness for cyclones as well as other support activities (such as education, reproductive health, self-help groups, and small and medium-sized enterprises), the number of women killed or affected by cyclones fell dramatically (IFRC-RCS 2002 in Sperling 2003). Finally, community-based work is not new in South and Southeast Asia. A number of community-based fisheries management or natural resource management projects have been implemented in these regions. Note, however, that strong involvement of local and national governments is required in implementing these types of initiatives.

Page 115: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

115

Box 5.7. Community-based adaptation to climate change in Viet Nam

This project was implemented in four communes and eight villages in Quang Dien and Phu Vang Districts, Thua Thien Hue province, in the north-central coast of Viet Nam in 2002. These villages experience about 30 days of flooding each year. In 1999, one of the worst floods resulted in the loss of hundreds of lives, along with property and other economic losses. This severe incident attracted international support for the Government of Viet Nam. During the relief operations, an initiative on “capacity building for adaptation to climate change” began. The main objective was to help build communities’ adaptive strategies for dealing with recurrent climatic catastrophes and to minimize the loss of lives and property. Three major steps are necessary for each participating community in this activity:

1. Scenario building includes identifying and analyzing the hazards, vulnerability to climate change, and existing and required adaptive capacity of the respective village. Interviews, focus group discussions, field surveys, historical profiling, and mapping of vulnerable sites are some methods applied to describe the current situation and future scenarios related to climate change. Adaptation mechanisms at the household and community levels, as well as social institutions that could contribute to hazard and disaster management strategies, are identified at this stage.

2. Planning involves discussions among the leaders of the social groups or organizations, such as those for farmers, fishers, women, youth, and other village political associations. Deliberations on threats and potential impacts arising from climate change and possible measures to address these issues are carried out at this stage. These measures can be livelihood improvements in agriculture and aquaculture, disaster management protocols, and other strategies. Participation of local government officials is critical during this process to ensure acceptance and implementation of the plan at the commune and district levels. In addition, the likelihood that the government will co-fund some subprojects identified at this point is higher. The main output at this stage is a “safer village plan” that will increase the resiliency of the community to the negative impacts of climate change.

3. Project implementation of some subprojects identified in the plan is made possible through in-kind and cash contributions to the community adaptation funds. These subprojects involve measures to ensure the safety of the people, infrastructure, and livelihoods of the village. Construction of an intercommune road, multipurpose school (as an emergency shelter), and technical support for agriculture and fisheries are provided. Training on the use of early warning devices and rescue and relief operations is extended to representatives of various social groups. Equipment such as boats, life jackets, and megaphones critical in giving timely warnings of impending disasters are made available to representatives of the social groups.

Source: Francisco 2008.

Government institutions

Government institutions play a significant role in ensuring the safety of the public, particularly during extreme natural disasters such as flooding. Discussions between government organizations and civil service institutions are important in identifying and implementing adaptation strategies for climate change. The participation of the government from the scenario-building process through the planning and implementation of adaptation activities is crucial. Furthermore, accountability of public institutions to the local society ensures good governance through responsive, participative, and accountable actions (Sperling 2003). Box 5.8 describes an example of public accountability in Bangladesh.

Page 116: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

116

Box 5.8. Government accountability related to flooding in Bangladesh

Flooding in Bangladesh is an annual incident, with one-third of cultivated land flooded in a normal monsoon year. In the northeastern part of the country, communities living near the Haor Basin learned to cope with flooding. Haor Basin was considered one of the most productive fishery resources in the floodplains, along with a food surplus that provided 10 percent of national grain supplies. The food system was unstable, however, and food shortages thus affect the communities, where 80 percent of workers are sharecroppers or landless laborers and a powerful elite controls the land and fishing rights. Although expected floods are manageable, flash floods can cause severe damage to homes and crops.

As a response to this threat, the Water Development Board built more than 800 kilometers of embankments. The government and communities are responsible for maintaining these embankments. Despite the flood control, a flash flood that hit the communities in 2002 damaged one-third of the embankment and 20 percent of crops, resulting in food shortages affecting an estimated 1.4 million people. Since then, the communities have complained about the lack of repairs to embankments, construction mismanagement, lack of monitoring, and corruption. As a result, the state minister for disaster management announced for the first time that elected officials will be engaged in embankment construction and maintenance. Since the local government lacks adequate capacity, a local NGO (HUNO) is working with the local government and the Water Development Board to develop a citizen-based monitoring system.

Source: Sashankar 2002; DFID 2002 in Sperling 2003.

Furthermore, financial support is necessary to implement any adaptation activities. Governments must be creative in finding ways to support activities with a limited budget. In addition, they must encourage collaboration with the private sector in developing climate change adaptation schemes, such as weather insurance.

Development agencies and donors

The core programs of international development agencies and donors must encompass the impacts of climate change as it affects poverty, food security, and economic development in developing countries. Development agencies must ensure that climate issues are internalized in their poverty reduction programs. This approach requires developing tools and methodologies, training, and raising awareness of senior management and staff. It may also involve modifying their own institutional processes to assure that climate change vulnerability in developing countries is addressed in all of their development work (Sperling 2003).

Although funds for climate change adaptation and mitigation strategies for developing countries are already available (through, for example, the UNFCC, ADB’s climate change funds, and other funding from regional and multilateral agencies), getting access to these funds poses a challenge to developing countries.

Private sector

Risk sharing or risk transfer is critical in implementing adaptation measures. Weather insurance markets normally developed by the private sector are a form of risk transfer (Francisco 2008). The insurance and reinsurance industries in particular have started to engage in adaptation activities in developing countries focusing on developing new risk-transfer products such as microinsurance, weather and crop insurance, and other mechanisms such as risk pooling and disaster-related bonds.

Page 117: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

117

Another important role for the private sector lies in development of new breeds and cultivars as has been described above.

5.7 Adaptation policy recommendations

Adaptation measures should be context and project specific. Criteria to consider include net economic benefits; timing of benefits; distribution of benefits; consistency with development objectives; consistency with other government policy costs; environmental impacts; spillover effects; implementation capacity; and social, economic, and technical barriers (Leary et al. 2007). Once the adaptation strategy has been evaluated, the measure that yields the greatest net benefit should be chosen. Methods presented by Fankhauser (1997), Callaway, Ringius, and Ness (1999), and Callaway (2003) have been integral in developing the cost–benefit analysis of adaptation strategies. The technical capability to change or improve agricultural practices can be assessed by determining their agronomic potential. Therefore, multiple criteria should be used to make judicious selections of adaptation measures from environmental, technical, social, and economic standpoints.

Adaptation to climate change in the Asia-Pacific region will require advancements in three areas: (1) pursuing innovative adaptation to climate change, (2) strengthening important ongoing development initiatives to support climate change adaptation, and (3) implementing climate change adaptation investments and policies. For all countries in the Asia and Pacific region depending on agriculture, a focus on agricultural productivity enhancement will be the key to adaptation to climate change.

Given the shifts in volume of rainfall and increased temperatures for the Asia Pacific region, as well as sea-level water rise and glacier melt, investments focusing on enhanced water control, water management and efficiency will be crucial for adaptation to climate change, particularly in Bangladesh, India, Viet Nam, Nepal, Bhutan, and the Pacific Islands. Knowledge and information sharing among farmers, government implementing agencies, and researchers should be given an enabling environment that supports adaptive management. Crop breeding will be an essential component of adapting to key biotic and abiotic stresses related to climate change, including drought, heat, salinity, pests, and disease. Biotechnology and genetic modification will be an increasingly large component of crop breeding because of the nature of upcoming climate change stresses.

Page 118: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

118

CHAPTER 6. CONCLUSIONS AND POLICY RECOMMENDATIONS

This report has reviewed the state of knowledge on climate change and agriculture in Asia and to some extent the Pacific Islands, as well as ongoing coping mechanisms and potential adaptation strategies. This chapter briefly synthesizes the salient findings in Chapters 2–5 discussing the severity of the impacts of climate change on agriculture, the impacts on food security, agriculture as a contributor to GHG emissions, adaptation and mitigation measures, and the various actors that have critical roles in mainstreaming and implementing climate change policies in the agricultural sectors in developing countries of Asia and the Pacific Islands. Based on our results, we identify seven key messages for the Asia and Pacific region:

Climate change will have significant impacts on agriculture in the Asia and Pacific region. Given the role of agriculture in employment, economic development, and global food security, adverse impacts on agriculture are of particular concern for the Asia and Pacific region.

Sound development policies are necessary but not sufficient for adaptation of agriculture to climate change in the Asia and Pacific region and elsewhere. A pro-growth, pro-poor development agenda that supports agricultural sustainability, including more targeted assistance, improves resilience and climate change adaptation. Adaptation to climate change will also, however, require targeted investments in agriculture.

Targeted assistance should be directed at those countries most vulnerable to climate change in the Asia and Pacific region—that is, those with large exposure to climate change impacts, high sensitivity to the impacts from climate change, and low adaptive capacity. Countries particularly vulnerable to climate change include Afghanistan, Bangladesh, Cambodia, India, Lao PDR, Myanmar and Nepal —with poor outcomes in all three vulnerability components—revealing South and Southeast Asia as the region most vulnerable to climate change in the Asia and Pacific region.

International agricultural trade is an important mechanism for sharing climate change risk. A more open global trading regime would increase resilience to climate change’s impacts.

Agriculture can help mitigate greenhouse gas emissions in Asia and the Pacific with appropriate incentive mechanisms and innovative institutions, technologies, and management systems. Incorporation of agricultural adaptation and mitigation in the ongoing international climate change negotiations must happen now in order to open opportunities for financing of sustainable growth under climate change. Mitigation strategies that support adaptation should be favored.

Cooperation among governments in Asia and the Pacific is necessary to ensure effective implementation of adaptation and mitigation strategies in their respective countries, as well as to explore financial means for addressing climate change. Funding modalities related to climate change (and accessibility of these funds to the vulnerable people), such as the Clean Development Mechanism, payment for environmental services, or other mechanisms to mitigate GHGs, must be implemented by Asian development planners and policymakers. Climate action plans need to be integrated into Poverty Reduction Strategy Plans or other national

Page 119: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

119

development plans. Without this integration, climate adaptation plans may simply add another layer of planning rather than aiding the mainstreaming process. Actors at all levels are called to action in the effort to adapt to climate change.

Uncertainty about where climate changes will have impacts must be reduced through more spatial analysis, as well as improved information, generated by both local agencies and users and science, which will then be disseminated to users.

These messages are discussed in more detail in the following sections.

Climate change will have significant impacts on agriculture in the Asia and Pacific region. Given the role of agriculture in employment, economic development, and global food security, adverse impacts on agriculture are of particular concern for the region.

Climate change hinders development in all sectors, not only in Asia, but globally. It is negatively affecting agriculture, particularly by intensifying the struggle for land and water resources. Understanding the adverse impacts on agriculture in the Asia and Pacific region is of particular importance because agriculture plays a crucial role in ensuring inclusive and sustainable development and because agricultural growth contributes to the attainment of the Millennium Development Goals, particularly the goals on hunger, poverty, environmental sustainability, water access, and to some extent, health.

Agriculture is the principal source of livelihood for more than 60 percent of the population of the Asia and Pacific region. Moreover, agriculture is the sector most vulnerable to climate change. Therefore, the effects on food production systems will directly affect the primary source of income for billions of people in the region, and perturbations in the food supply will have overall implications for the wider population of net food purchasers. Finally, the Asia and Pacific region accounts for half of global cereal supply and demand. Any significant changes in the food systems of this region will have implications for food supply and food prices globally.

Climate change will have significant negative impacts on agricultural production, with the greatest reductions occurring in developing countries of Asia. Results of the modeling scenarios discussed in Chapter 2 of this report consistently show that South Asia and Southeast Asia will be severely affected by climate change, threatening food security in these subregions. In contrast, parts of East Asia may find their agricultural capacity enhanced.

Serious declines in agricultural production will adversely affect agricultural GDP in many Asian countries, and grave climatic conditions will cause heavy economic losses in Pacific Island countries. The projected decline in production due to climate change will likely lead to substantial increases in food prices. While these predictions have been shown across a number of models, specific effects will differ by region. The effects of multiple stresses, such as extreme weather events, pests, and diseases, have not been adequately considered and require additional analysis.

Page 120: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

120

Sound development policies are necessary but not sufficient for adaptation of agriculture to climate change in the Asia and Pacific region and elsewhere. A pro-growth, pro-poor development agenda that supports agricultural sustainability, including more targeted assistance, improves resilience and climate change adaptation. Adaptation to climate change will, however, also require targeted investments in agriculture.

The costs of agricultural adaptation to climate change in Asia will be significant, beyond the investments in agricultural research, water, rural roads, and social services that are already required to meet development and poverty reduction goals.

Climate change can become the stimulus for implementing difficult but necessary changes. Managing climate change as an international public good creates opportunities for new markets and pricing policies that can help meet longer-term goals of sustainability through the valuation of resources. Rising prices of carbon, food, fuel, and environmental resources due to climate change could stimulate significant policy and investment opportunities. Instead of seeing climate change as a tax on growth, countries can benefit by implementing low-cost, resource-conserving technologies in the agricultural sector. They can exploit synergies between building ecosystem resilience and agricultural productivity through a focus on agricultural productivity enhancement, and new agricultural financing mechanisms such as carbon markets.

A pro-growth, pro-poor development agenda that supports agricultural sustainability, including more targeted foreign assistance in Asia, also improves resilience and climate change adaptation. A broad set of technical skills will be needed to plan for and respond to a wide range of unpredictable contingencies, and the backbone of these efforts will be improved knowledge, coordination, collaboration, information exchange, and institutional responsiveness. Building resilience—especially among the poor—will require enhancing the adaptive capacity of individuals and institutions to deal with uncertainties in their local settings through the testing and scaling up of effective pilot projects.

Effective adaptation requires more than good development policy, however, for a number of reasons. First, because climate change has a negative impact on agricultural production in most developing countries, achieving any given food security target will require greater investments in agricultural productivity. Key areas for increased investment include agricultural research, irrigation, rural roads, information technologies, market support, and extension services.

Second, the allocation of investment within and across sectors will need to change to achieve effective adaptation. Investments in agricultural research will need to shift to focus on traits relevant to climate change adaptation, such as drought and heat tolerance, insect and pest tolerance, and nitrogen use efficiency, which can reduce carbon emissions while promoting agricultural productivity. Biotechnology and genetic modification will be an increasingly large part of crop breeding because of the need for wider genetic variation to adapt to climate change stresses. In irrigation and water resources, investments may be needed to expand large-scale storage to deal with the increased variability of rainfall and runoff. On the other hand, in regions where changes in precipitation are highly uncertain, investments might better be distributed in a variety of small catchments. Climate change and variability in water supply, together with potential long-term changes in the cost of energy, could also dramatically change the cost-benefit calculus for big dams for storage, irrigation, and hydropower, making these investments more

Page 121: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

121

attractive despite the environmental and human relocation issues that dams raise. The appropriate level and location of future irrigation investments could also change dramatically.

A third way that adaptation is different from business-as-usual development policy is that greater variability in weather and production outcomes will require enhanced attention to risk-sharing and risk-reducing investments. Such investments include financial market innovations, climate-based crop insurance, and broad-based social safety nets, which both protect against the negative impacts of increased risk and induce farmers to make decisions that are not too risk-averse. International agricultural trade is an important mechanism for sharing climate change risk, so open trading regimes should be supported. Appropriate agricultural advisory services, hydro-meteorological infrastructure, functioning financial markets, and effective institutions are necessary to minimize the risks to farmers as they make decisions about agricultural production. Also directly related to managing risk is the need to upgrade the efficiency and sophistication of infrastructure and other investments, including modernizing instead of rehabilitating irrigation and investing in paved, not dirt, roads. More sophisticated agricultural practices, such as integrated pest management, are also needed, requiring improvement in human capacity in agricultural management. Strengthening women’s roles in household and agriculture production, as well as their rights and control of assets, would improve the effectiveness of risk management.

The fourth area is the spatial location of investments. Investments should be targeted to regions where the benefits become higher because of climate change and should be reduced in areas where climate change is so severe that production is no longer feasible. Sea-level rise will increase the high concentration of salt in farm areas, which may require retooling of production systems. In some areas, for example, instead of producing crops, farmers may be better off pursuing alternative livelihoods, such as raising livestock, as practiced in the southwestern coastal areas of Bangladesh during flooding season. More and better spatial analysis is needed to reduce uncertainty about where climate change will have impacts.

A fifth way in which climate change is a game changer is that the costs of bad policies are made worse by climate change, which will contribute to increased food, energy, and water prices. Perverse subsidies for water, energy, and fertilizer should be removed, with the savings invested in adaptation activities that boost farm income. These subsidies have not only distorted production decisions, but also encouraged carbon emissions beyond economically appropriate levels. As the real prices of natural resources rise, market-based approaches for managing environmental services in response to climate change (such as through water pricing, payment for environmental services, and carbon trading) will be increasingly important. Improved definition and protection of land and water property rights will be necessary to effectively implement market-based approaches to climate change policy.

The sixth difference is that the recognition of carbon as a global externality and the valuation of carbon through carbon trade increase the value of sustainable farming practices. This situation improves the likelihood that farmers will adopt long-term sustainable farming practices such as minimum tillage; integrated soil fertility management; and integrated pest, disease, and weed management.

Implementation of these adaptation and mitigation measures can be realized only through enhanced agricultural investments. There is a strong need to revisit investment priorities and opportunities at the national level in Asia and the Pacific. Developing countries have chronically underinvested in science, technology, and innovation. Growth in public investments in research

Page 122: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

122

stagnated after the 1980s in developing countries. Investments in biotechnology and biosafety, especially by the public sector, may be insufficient to address pressing needs in both areas and to resolve national constraints. In spite of the limitations, the public sector in many developing countries has invested in agricultural biotechnology research, yet few of its technologies have reached the commercialization stage. Many developing countries, particularly those in Southeast Asia, need to develop the minimal infrastructure and scientific capacity to master and implement risk assessments and biosafety regulations. Investments in biotechnology, including GM crops, could provide a transformational approach to addressing the trade-offs between energy efficiency and agricultural productivity. Biotechnology could profoundly affect future demand for freshwater and investment requirements in irrigation and other water sectors. GM crops have the potential to address major water-related stresses under both rainfed and irrigated farming and possibly to offer solutions to important water-quality problems. Breeding crop varieties with high water-use efficiency—a good indicator of the crop’s ability to withstand environmental stresses, particularly drought and salinity—is thus one policy option. Biotechnology’s role as a possible substitute for large-scale water investments must be considered in future planning for irrigation and water supply and sanitation investments.

Increased and diversified investments are needed in plant breeding, livestock improvement, and other interventions at the biological and molecular levels to enhance agricultural productivity in ways that ultimately contribute to poverty reduction, agricultural development, and economy-wide growth throughout the region. Such a program requires heavy investment in advanced scientific expertise and equipment, as well as a political and social commitment to long-term funding of agricultural science and technology at levels significantly greater than current funding. Furthermore, it requires new investments to create organizations that are more dynamic, responsive, and competitive than the public organizations that make up the bulk of national agricultural research and extension systems in Asia today.

Major investments in water infrastructure are also needed. Dams have proven to be an effective way to protect agricultural systems and human settlements from water variability, and a higher demand for dams is expected as a result of increasing water variability and energy demand. Big dams are known, however, to cause considerable environmental and social impacts. Furthermore, investment is needed in engineering techniques to reduce environmental impacts, management techniques to optimize their use, planning tools to reduce social impacts, and tools to improve design and operational techniques. Investments should also be made to scale up underground storage techniques. Finally, more investments should be made in research on the viability of inter-basin transfer schemes, which can be politically challenging and risky in light of future uncertainty about water availability.

Policies that favor private sector investment in crop improvements targeted to climate change in the developed and developing world are critical. These policies include (1) decreasing the bureaucratic hurdles to business formation; (2) developing infrastructure that enables production and distribution of improved seeds and other agricultural inputs; (3) developing appropriate regulatory and biosafety protocols for introduction of transgenic cultivars; and (4) reforming intellectual property rights that could encourage private investment in crop improvement.

Meeting the challenges of climate change adaptation in agriculture requires long-term investment by farmers. Long-term investment (in areas such as integrated soil fertility management, tree planting, and water harvesting) in turn requires secure property rights to

Page 123: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

123

provide people with the incentive and authority to make the investments. By changing the profitability of land, such as through the potential for income from carbon markets and biofuels, climate change may worsen the position of farmers with insecure property rights, leading to expulsion from their land as landlords seek to increase their share of the new income streams. Improvement in land rights is therefore an essential component in effective and equitable adaptation.

Targeted assistance should be directed at those countries most vulnerable to climate change in the Asia and Pacific region—that is, those with large exposure to climate change impacts, high sensitivity to the impacts from climate change, and low adaptive capacity. These countries include Afghanistan, Bangladesh, Cambodia, India, Lao PDR, Myanmar and Nepal—with poor outcomes in all three vulnerability components—revealing South and Southeast Asia as the region most vulnerable to climate change in the Asia and Pacific region.

Vulnerability to climate change depends not only on exposure to climate events, but also on physical, environmental, socioeconomic, and political factors that influence how sensitive countries will be to a changing climate and how they will be able to cope. Climate change is likely to increase the vulnerability of poor farmers who already struggle with land degradation.

Countries in South Asia, Southeast Asia, and the Pacific Islands are highly vulnerable to sea-level rise, which will increase the risk of floods. Glaciers in the Himalayas and Central Asia are already melting as a result of global warming. In regions highly dependent on livestock production, such as Mongolia and Inner Mongolia, China, overgrazing increases vulnerability to climate change. Afghanistan, Bangladesh, India, and Nepal are particularly vulnerable to climate change, and South Asia is the region most vulnerable to climate change in the Asia and Pacific region.

Vulnerability assessments are important to ensure that scarce public and private resources are allocated to those most in need of adapting to climate change. Although various vulnerability assessments generally come to similar conclusions, differences in results do exist because of the use of different data, different factors representing vulnerability, and differing methodologies. Care must therefore be taken when drawing further conclusions or basing investment decisions on such assessments.

As shown by Bangladesh’s improved resiliency to tropical cyclones between 1991 and 1997, adaptation is possible even for the most destitute and vulnerable countries.

International agricultural trade is an important mechanism for sharing climate change risk. A more open global trading regime would increase resilience to climate change’s impacts.

Rule-based, fair, and free international trade is particularly critical in times of crisis, as the export ban problems following the food price crisis of 2007–08 underline. A sound global trading system is especially crucial in the context of climate change. As shown in Chapter 2, the impacts of climate change on agricultural growth and production will likely make many developing countries increasingly reliant on food imports. To increase confidence in international agricultural trade, the WTO Doha Round should be completed; OECD countries should reduce

Page 124: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

124

or eliminate trade restrictions that limit developing-country export access to markets; and buffering mechanisms should be established to better address volatility in world markets. Alternative or complementary approaches to market stabilization for cereals include a joint pooling of fixed portions of national stocks into an international grain reserve and/or a financial facility, provided by the International Monetary Fund (IMF), for imports by countries during food emergencies

Agriculture can help mitigate greenhouse gas emissions in Asia and the Pacific with appropriate incentive mechanisms and innovative institutions, technologies, and management systems. Incorporation of agricultural adaptation and mitigation in the ongoing international climate change negotiations must happen now in order to open opportunities for financing of sustainable growth under climate change. Mitigation strategies that support adaptation should be favored.

In general, agriculture affects climate change significantly through livestock production and the conversion of forest to land cover that has low carbon sink or sequestration potential. Nitrous oxide emissions from crop production and methane from rice production are also significant contributors.

Because agriculture is still the main livelihood source for more than half of the people in the Asia and Pacific region, benefit streams for Asia from mitigation strategies have the potential to contribute to poverty reduction, food security, and the resilience of agroecological systems. Small farmers have significant potential to sequester soil carbon if appropriate policy reforms are implemented. Successful implementation of soil carbon trading would generate significant co-benefits for soil fertility and long-term agricultural productivity. If the high transaction costs for small-scale projects can be eliminated, carbon markets could be a significant source of financing. The benefit stream from mitigation of 276.79 Mt CO2-eq a year at a carbon price of US$20/t CO2-eq in agriculture could amount to US$5.5 billion annually for Asia, accounting for 18 percent of total global mitigation potential.

The use of high-yielding varieties, a shift to rice-wheat production systems, and alternating dry-wet irrigation are technologies that combine mitigation and adaptation objectives through reducing emissions, conserving water, and reducing land requirements and fossil fuel use. Other mitigation strategies that have substantial synergic effects with adaptation and food security for rural communities in Asia and the Pacific include the restoration of degraded soils and efficient water use in crop cultivation. All of these strategies help conserve soil and water resources and at the same time enhance ecosystem functioning, including water use efficiency and crop resiliency against pests, diseases, and extreme climate events. GHG emissions from agriculture can be further mitigated through nutrient, water, and tillage management; improved crop varieties (particularly rice, the main staple in Asia); and use of crop residues for renewable energy and carbon sequestration. Improved pasture management to control livestock overgrazing will help decelerate desertification.

Although there are viable mitigation technologies in the agricultural sector, key constraints need to be overcome. First, the rules of access—which still do not credit developing countries for reducing emissions by avoiding deforestation or improving soil carbon sequestration—must change. And second, the operational rules, with their high transaction costs for developing countries, and small farmers and foresters in particular, must be streamlined.

Page 125: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

125

Policies focused on mitigating GHG emissions, if carefully designed, can help create a new development strategy that encourages the creation of more valuable pro-poor investments by increasing the profitability of environmentally sustainable practices. To achieve this goal, it will be necessary to streamline the measurement and enforcement of offsets, financial flows, and carbon credits for investors. It is important to enhance global financial facilities and governance to simplify rules and increase funding flows for mitigation in developing countries.

Cooperation among governments in Asia and the Pacific is necessary to ensure effective implementation of adaptation and mitigation strategies in their respective countries, as well as to explore financial means for addressing climate change. Funding modalities related to climate change, such as the Clean Development Mechanism, payment for environmental services, and other mechanisms to mitigate GHGs, must be implemented by Asian development planners and policymakers. Climate action plans need to be integrated into Poverty Reduction Strategy Plans or other national development plans. Without this integration, climate adaptation plans may simply add another layer of planning rather than aiding the mainstreaming process. Actors at all levels are called to action in the effort to adapt to climate change.

Adaptation to climate change, typically treated as a stand-alone activity, should be integrated into development projects, plans, policies, and strategies. Development policy issues must inform the work of the climate change community, and development and climate change perspectives should be combined in integrated approaches that recognize how persistent poverty and environmental needs exacerbate the adverse consequences of climate change. Climate change will alter the set of appropriate investments and policies over time, both in type and in spatial location. Effective adaptation therefore requires that policymakers not only judiciously select measures within their policy context and strategic development framework, but also explicitly target the impacts of climate change, particularly on the poor.

Development policies and plans at all levels need to consider the impacts of climate change on the agricultural sector. National and regional policymakers must integrate the effects of climate change and the outcomes of assessments and scenarios into their national agricultural plans and policies. Moreover, mainstreaming should aim to limit development policies and plans that inadvertently encourage, rather than minimize, vulnerability to the impacts of climate change.

Achieving these goals will require the engagement of a core ministry, such as the Ministry of Finance or Planning, alongside the Ministry of Agriculture, to ensure strong government support. Second, the core capacities of developing-country governments will need to be further developed. Such capacity building is required in a number of areas, including technical subjects such as climate forecasting and scenario planning, as well as general development topics such as governance, accountability, and empowerment of local communities. Third, adaptive and flexible management will be essential to address the broadening nature and increasing severity of potential climate impacts in a given area and the unavoidable uncertainties associated with predicting these impacts.

Page 126: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

126

New mechanisms to support adaptation, including the Least Developed Country Fund, the Special Climate Change Fund, and the Adaptation Fund, provide the opportunity to mainstream adaptation into local and regional development activities.

Short-term regional adaptation initiatives could include transboundary or regional adaptation evaluation exercises and investment assessments. Medium-term regional adaptation initiatives could include the development of agricultural climate information systems, regional disaster and emergency relief funds, and larger-scale infrastructure development. Regional initiatives should be supported by climate change interest groups staffed by experts from the Ministries of Agriculture and Finance of the Asia and Pacific region.

In addition, the private sector—the insurance and reinsurance industries in particular—needs to engage more in adaptation activities in developing countries, building on the risk-transfer products they have already started to develop, such as microinsurance, weather and crop insurance, and disaster-related bonds.

Uncertainty about where climate change will have impacts must be reduced through more spatial analysis, as well as improved information, generated by both local agencies and users and science, which will then be disseminated to users.

Significant uncertainty remains regarding climate change impacts on agriculture (and other sectors). Disagreements among modeling studies with regard to the future impacts of climate change on agricultural capacity and crop yields are in part a result of different assumptions. Another major limitation is a lack of integrated assessment incorporating all key climate variables. Many climate variables have feedback effects among themselves, which are left out of already complex modeling efforts. Furthermore, almost all climate change modeling efforts leave out several key factors. Climate extreme events and other stressors—such as increased climate variability, sea-level rise, and land degradation—are often partially or entirely ignored. In agriculture, pest and disease aspects and their feedback effects are seldom taken into account. Another major limitation is that the quality and extent of research vary by country; much information is available on China, and little if any is available for Central Asia and the Pacific Island states.

Page 127: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

127

REFERENCE LIST

———. 1997. Review of the state of world fishery resources: Marine fisheries. FAO Fisheries Circular No. 920 FIRM/C920. Rome.

———. 2004. Agricultural employment. Rome.

———. 2005. World development indicators. Washington, DC.

———. 2006. The role of markets and governments in helping society adapt to a changing climate. Climatic Change 78 (1): 203–15.

———. 2006. The state of food insecurity in the world. Rome.

———. 2006. The State of Food Insecurity in the World. Rome: FAO

———. 2007. FAO statistical yearbook: Countries profile—Uzbekistan. Rome.

———. 2007. Food security and climate change in the South Pacific. Pacific Ecologist (Winter): 32–36.

———. 2007a. Socio-economic-techno-environmental assessment of IDEI products (Treadle pump). Project Report No. 2006RR24. New Delhi. Unpublished.

———. 2007b. Socio-economic-techno-environmental assessment of IDE products (Low cost KB Drip). Project Report No. 2006RR24. New Delhi. Unpublished.

———. 2008. Climate change and food security in Pacific Island countries. Rome.

———. 2008. Climate change and food security: A framework document. Rome.

———. 2008. Protecting health from climate change: World Health Day 2008. Geneva.

———. 2008. World development indicators. Washington, DC.

———. 2008a. Climate change and food security in Pacific Island Countries. Rome.

———. 2008b. Climate change and food security: A framework document. Rome.

———. 2008b. Food preservation and canning in mountain communities of central Tajikistan. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=178.

———. 2008c. Climate change, bioenergy, and land tenure: Technical background document. Rome.

———. 2008c. Dam-fields in northwest China. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=40.

———. 2008d. Facts and figures. http://www.fao.org/Gender/en/lab-e.htm (accessed December 30, 2008).

———. 2008d. Post-flood rehabilitation in Bangladesh. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=160.

Page 128: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

128

———. 2008e. Early or late transplanted aman rice production in Bangladesh. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=194.

———. 2008e. The state of food insecurity in the world. Rome.

———. 2008f. Hydroponics in Bangladesh. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=80.

———. 2008g. Spate irrigation in Sidh, Pakistan. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=69.

———. 2008h. Kunds in Thar Desert, India. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=57.

———. 2008i. Drip irrigation in Bhutan. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=21.

———. 2008j. Diversifying crops in Arunachal Pradesh, India. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=79.

———. 2008k. Homestead gardens in Bangladesh. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=192.

———. 2008l. Indigenous forecasting in western Himalayas. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=46.

———. 2009. Enabling agriculture to contribute to climate change mitigation. An FAO submission to the UNFCCC Ad Hoc Working Group on Long-Term Cooperative Action. Rome.

Adger, W. N. 2001. Social capital and climate change. Tyndall Centre for Climate Change Research Working Paper No. 8. Norwich, UK: Tyndall Centre for Climate Change Research.

Adger, W. N., N. W. Arnella, and E. L. Tompkins. 2005. Successful adaptation to climate change across scales. Global Environmental Change 15 (2): 77–86.

ADPC (Asian Disaster Preparedness Center). 2003. Climate change and development in Viet Nam: Agriculture and adaptation for the Mekong Delta region. Eschborn: Deutsche Gesellschaft für Technische Zusammenarbeit (GTZ).

Aggarwal, P. K., and R. K. Mall. 2002. Climate change and rice yields in diverse agro-environments of India. II. Effect of uncertainties in scenarios and crop models on impact assessment. Climatic Change 52 (3): 331–43.

Alderman, H., J. Hoddinott, and B. Kinsey. 2003. Long-term consequences of early childhood malnutrition. FCND Discussion Paper No. 168. Washington, DC: Internationa Food Policy Research Institute.

Ali, A. 1999. Climate change impacts and adaptation assessment in Bangladesh. Climate Research 12 (2–3): 109–16.

Page 129: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

129

Allouche, J. 2004. A source of regional tension in Central Asia: The case of water. In The illusions of transition: Which perspectives for Central Asia and the Caucasus? Geneva: CIMERA.

Angerer, J., G. Han, I. Fujisaki, and K. Havstad. 2008. Climate change and ecosystems of Asia with emphasis on Inner Mongolia and Mongolia. Rangelands 30 (3): 46–51.

Ansorg, T., and T. Donelly. 2008. Climate change in Bangladesh: Coping and conflict. European Security Review No. 40. Brussels: International Security Information Service (ISIS) Europe.

Arnell, N. W. 2003. Effects of IPCC SRES emissions scenarios on river runoff: A global perspective. Hydrology and Earth System Sciences 7 (5): 619–41.

Arnell, N. W. 2004. Climate change and global water resources: SRES emissions and socio-economic scenarios. Global Environmental Change-Human and Policy Dimensions 14 (1): 31–52.

Arnell, N. W., M. G. R. Cannell, M. Hulme, R. S. Kovats, J. F. B. Mitchell, R. J. Nicholls, M. L. Parry, M. T. J. Livermore, and A. White. 2002. The consequences of CO2 stabilisation for the impacts of climate change. Climatic Change 53 (4): 413–46.

Baker, J. M., T. E. Ochsner, R. T. Venterea, and T. J. Griffis. 2007. Tillage and soil carbon sequestration—What do we really know? Agriculture, Ecosystems and Environment 118 (1–4): 1–5.

Banerjee, L. 2007. Flood disasters and agricultural wages in Bangladesh. Development and Change 38 (4): 641–64.

Bannon, I., and S. Naraghi-Anderlini. 2009. Module 11: Gender and crises: Implications for agriculture. In Gender in agriculture sourcebook. Washington, DC: World Bank, Food and Agriculture Organization of the United Nations, International Fund for Agricultural Development.

Baradas, M. W., and J.G. Mina. 1996. Water management 2000: Not by irrigation alone. Presented at Workshop on Food Security, Los Baños, Laguna, Philippines.

Barnett, J. 2003. Security and climate change. Global Environmental Change-Human and Policy Dimensions 13 (1): 7–17.

Barnett, J., and W. Adger. 2003. Climate dangers and atoll countries. Climatic Change 61 (3): 321-337

Barnett, J., and W. Adger. 2003. Climate dangers and atoll countries. Climatic Change 61 (3): 321–37.

Bates, B., Kundzewick, Z. W., Wu, S., and Palutikof, J., ed. 2008. Climate change and water. IPCC Technical Paper No. VI.

Bates, B., Z. W. Kundzewick, S. Wu, and J. Palutikof, ed. 2008. Climate change and water. Technical Paper of the Intergovernmental Panel on Climate Change, IPCC Secretariat, Geneva, 210 pp.

Page 130: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

130

Batima, P. 2006. Climate Change Vulnerability and Adaptation in the Livestock Sector of Mongolia. No. Project No. AS 06. Washington, DC: The International START Secretariat

Batima, P. 2006. Climate change vulnerability and adaptation in the livestock sector of Mongolia. Project No. AS 06. Washington, DC: International START Secretariat

Batima, P., L. Natsagdorj, N. Batnasan, and M. Erdenetuya. 2005. Mongolia's livestock system vulnerability to climate change. In For Whom the Bell Tolls, Case Studies of Climate Change Vulnerability. Ed. N. A. Leary, C. Conde, A. Nyong, and J. Pulhin

Bauer, A., R. Hasan, R. Magsombol, and G. Wan. 2008. The World Bank's New poverty data: Implications for the Asian Development Bank. ADB Sustainable Development Working Paper Series No. 2. Manila: Asian Development Bank

Bindoff, N. L., J. Willebrand, V. Artale, A. Cazenave, J. Gregory, S. Gulev, K. Hanawa, C. Le quere, S. Levitus, Y. Nojiri, C. K. Shum, L. D. Talley, and A. Unnikirshnan. 2007. Observations: Oceanic climate change and sea level. In Climate change 2007: The physical science basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, and H. L. Miller. Cambridge, UK: Cambridge University Press.

Binns, P. 2007. Sowing the seeds of prosperity: Developing bioenergy technology to alleviate smallholder farmer poverty. Proposal to the Bill and Melinda Gates Foundation. Unpublished. Westbrook Associates, LLC, Seattle.

Black, R. E., L. H. Allen, Z. A. Bhutta, L. E. Caulfield, M. de Onis, M. Ezzati, C. Mathers, and J. Rivera. 2008. Maternal and child undernutrition. 1 - Maternal and child undernutrition: Global and regional exposures and health consequences. Lancet 371 (9608): 243–60.

Boer, R. 2009. Impacts of climate change in agriculture. Country case study submitted to International Food Policy Research Institute, Washington, DC. Unpublished report.

Bolortsetseg, B., and G. Tuvaansuren. 1996. The potential impacts of climate change on pasture and cattle production in Mongolia. Water Air and Soil Pollution 92 (1–2): 95–105.

Bouma, M. J., C. Dye, and H. J. van der Kaay. 1996. Falciparum malaria and climate change in the Northwest Frontier Province of Pakistan. American Journal of Tropical Medicine and Hygiene 55 (2): 131–37.

Bouma, M., and H. van der Kaay. 1996. The El Niño Southern Oscillation and the historic malaria epidemics on the Indian subcontinent and Sri Lanka: An early warning system for future epidemics? Tropical Medicine and International Health 1 (1): 86–96.

Brody, A., J. Demetriades, and E. Esplen. 2008. Gender and climate change: Mapping the linkages: A scoping study on knowledge and gaps. Brighton , UK: BRIDGE, Institute of Development Studies.

Brooks, N., W. N. Adger, and P. M. Kelly. 2005. The determinants of vulnerability and adaptive capacity at the national level and the implications for adaptation. Global Environmental Change-Human and Policy Dimensions 15 (2): 151–63.

Page 131: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

131

Bryan, E., W. Akpalu, M. Yesuf, and C. Ringler. 2008. Global carbon markets: Are there opportunities for Sub-Saharan Africa? IFPRI Discussion Paper 832. Washington, DC: International Food Policy Research Institute.

Bryan, E., W. Akpalu, M. Yesuf, and C. Ringler. 2008. Global carbon markets: Are there opportunities for Sub-Saharan Africa? IFPRI Discussion Paper 832. Washington, DC: International Food Policy Research Institute.

Bryan, E., W. Akpalu, M. Yesuf, and C. Ringler. 2008. Global carbon markets: Are there opportunities for Sub-Saharan Africa? IFPRI Discussion Paper 832. Washington, DC: International Food Policy Research Institute.

Bryla, E. and J. Syroka. 2007. Developing index-based insurance for agriculture in developing countries. Sustainable Development Innovation Briefs Issue 2 (March). United Nations Department of Economic and Social Affairs. http://esa.un.org/marrakechprocess/pdf/InnovationBriefs_no2.pdf.

Bumb, B., and C. Baanante. 1996. World trends in fertilizer use and projections to 2020. 2020 Brief 38. Washington, DC: International Food Policy Research Institute.

Burton, I. and B. Lim. 2005. Achieving Adequate Adaptation in Agriculture. Climate Change 70, 191-200.

Cai, X., D. Wang, T. Zhu, and C. Ringler. 2009. Assessing the regional variability of GCM simulations. Geophysical Research Letters, 36, L02706, doi:10.1029/2008GL036443.

Callaway, J.M. 2003. Adaptation benefits and costs—measurement and policy issues. Environment Directorate, Environment Policy Committee, OECD. ENV/EPOC/GSP(2003)10/FINAL, Paris, France.

Callaway, J.M., Ringius, L., Ness, L. 1999. Adaptation costs: a framework and methods. In: Christensen, J., Sathaye, J. (Eds.), Mitigation and Adaptation Cost Assessment Concepts, Methods and Appropriate Use. UNEP Collaborating Centre on Energy and Environment, Risf National Laboratory, Roskilde, DK.

Campbell-Lendrum, D. H., R. S. Kovats, A. J. McMichael, C. Corvalan, B. Menne, and A. Pruss-Ustun. 2005. The global burden of disease due to climate change: Quantifying the benefits of stabilization for human health. Met Office. http://www.stabilisation2005.com/39_Sari_Kovats.pdf (accessed March 4, 2009)

Cannon, T. 2002. Gender and climate hazards in Bangladesh. Gender and Development 10 (2): 45–50.

Carvajal-Escobar, Y., M. Quintero-Angel, and M. Garcia-Vargas. 2008. Women's role in adapting to climate change and variability. Advances in Geosciences 14: 277–80.

Cazelles, B., M. Chavez, A. J. McMichael, and S. Hales. 2005. Non-stationary influence of El Niño on the synchronous dengue epidemics in Thailand. Epidemiology 16 (5): S156–S157.

Chan, K. Y., D. P. Heenan, and H. B. So. 2003. Sequestration of carbon and changes in soil quality under conservation tillage on light-textured soils in Australia: A review. Australian Journal of Experimental Agriculture 43 (4): 325–44.

Page 132: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

132

Chinvanno, S. 2003. Information for sustainable development in light of climate change in Mekong River Basin. Southeast Asia START Regional Center, Bangkok, Thailand.

Chinvanno, S., S. Souvannalath, B. Lersupavithnapa, V. Kerdsuk, and N. T. H. Thuan. 2006. Climate risks and rice farming in the lower Mekong River countries. AIACC Working Papers No. 40. Washington, DC: Assessments of Impacts and Adaptations to Climate Change (AIACC).

Christensen, J. H., B. C. Hewitson, A. Busuioc, A. Chen, X. Gao, R. Jones, W. T. Kwon, R. Laprise, V. Magana, L. Mearns, C. Menendez, J. Raisaenes, A. Rinke, R. K. Kolli, A. Starr, and P. Whetton. 2007. Regional Climate Projections. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Ed. S. Solomon, Qin.D., M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, and H. L. Miller. Cambridge: Cambridge University Press.

Christensen, L., M. B. Coughenour, J. E. Ellis, and Z. Z. Chen. 2004. Vulnerability of the Asian typical steppe to grazing and climate change. Climatic Change 63 (3): 351–68.

Chullun, T., L. L. Tieszen, and D. Ojima. 1999. Land use impact on C4 plant cover of temperate East Asian grasslands. In NIES workshop on information bases and modeling for land-use and land-cover changes studies in East Asia, ed. K. Otsubo. Ibaraki: Center for Global Environmental Research

CIFOR (Center for International Forestry Research). 2002. Making forest carbon works for low-income producers. Infobrief 2. Jakarta. Available at http://www.cifor.cgiar.org/publications/pdf_files/infobrief/002-Infobrief.pdf.

Cline, W. 2007. Country-level agricultural impact estimates. In Global warming and agriculture: Impacts estimates by country. Washington, DC: Peterson Institute.

Cline, W. 2007. Country-Level Climate Projections. In Global Warming and Agriculture: Impacts Estimates by Country. Washington, DC: Peterson Institute.

Cline, W. 2007. Global warming and agriculture: Impact estimates by country. Washington, DC: Peterson Institute.

Cohen, M., C. Tirado, N. Aberman, and B. Thompson. 2008. Impact of climate change and bioenergy on nutrition. Washington, DC: International Food Policy Research Institute.

Confalonieri, U., B. Menne, R. Akhtar, K. L. Ebi, M. Hauengue, R. S. Kovats, B. Revich, and A. Woodwart. 2007. Human health. In Climate change 2007: Impacts, adaptation, and vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. M. Parry, O. F. Canziani, J. Palutikof, P. J. van der Linden, and C. E. Hanson. Cambridge, UK: Cambridge University Press.

Crisis Group. 2002. Central Asia: Water and conflict. Asia Report No. 34. Brussels: International Crisis Group.

Cruz, R. V. O., H. Harasawa, M. Lal, S. Wu, Y. Anokhin, B. Punsalmaa, Y. Honda, M. Jafari, C. Li, and N. Huu Ninh. 2007. Asia. In Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the

Page 133: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

133

Intergovernmental Panel on Climate Change. Ed. M. Parry, O. F. Canziani, J. Palutikof, P. J. van der Linden, and C. E. Hanson. Cambridge, UK: Cambridge University Press.

Cruz, R. V. O., H. Harasawa, M. Lal, S. Wu, Y. Anokhin, B. Punsalmaa, Y. Honda, M. Jafari, C. Li, and N. Huu Ninh. 2007. Asia. In Climate change 2007: Impacts, adaptation, and vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. M. Parry, O. F. Canziani, J. Palutikof, P. J. van der Linden, and C. E. Hanson. Cambridge, UK: Cambridge University Press.

Cruz, R. V. O., H. Harasawa, M. Lal, S. Wu, Y. Anokhin, B. Punsalmaa, Y. Honda, M. Jafari, C. Li, and N. Huu Ninh. 2007. Asia. In Climate change 2007: Impacts, adaptation, and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. M. Parry, O. F. Canziani, J. Palutikof, P. J. van der Linden, and C. E. Hanson. Cambridge, UK: Cambridge University Press.

Dankelman, I. 2002. Climate change: Learning from gender analysis and women's experiences of organising for sustainable development. In Gender, development, and climate change, ed. R. Masika. Oxford, UK: Oxfam.

Dankelman, I., K. Alam, W. B. Ahmed, Y. D. Gueye, N. Fatema, and R. Mensah-Kutin. 2008. Gender, climate change, and human security. Lessons from Bangladesh, Ghana, and Senegal. New York: Women's Environment and Development Organization.

Dasgupta, S., B. Laplante, C. Meisner, D. Wheeler, and J. Yan. 2007. The Impact of Sea Level Rise on Developing Countries: A Comparative Analysis. World Bank Policy Research Working Paper No. 4136. Washington, DC: World Bank

Delate, K., M. Duffy, C. Chase, A. Holste, H. Friedrich, and N. Wantate. 2003. An economic comparison of organic and conventional grain crops in long-term agro-ecological research (LTAR) site in Iowa. American Journal of Alternative Agriculture 18 (5): 59–69.

Delgado, C., M. Rosegrant, H. Steinfeld, S. Ehui, C. Courbois. 1999. Livestock to 2020: The next food revolution. 2020 Discussion Paper 28. Washington, DC: International Food Policy Research Institute.

DENR (Department of Environment and Natural Resources). 1999. The Philippines' initial national communication on climate change. Quezon City, Philippines.

Dey, M., and M. Prein. 2005. Increased income from seasonally flooded rice fields through community-based fish culture in Bangladesh and Viet Nam. Plant Production Science 8 (3): 349–53.

Dilley, M., R. S. Chen, U. Deichmann, A. L. Lerner-Lam, M. Arnold, J. Agwe, P. Buys, O. Kjekstad, B. Lyon, and G. Yetman. 2005. Natural disaster hotspots: A global risk analysis synthesis report. Washington, DC: World Bank.

Easterling, W. E., P. K. Aggarwal, P. Batima, K. M. Brander, E. D. Lin, S. M. Howden, A. P. Kirilenko, J. Morton, J. Soussana, J. Schmidhuber, and F. Tubiello. 2007. Food, fibre and forest products. In Climate change 2007: Impacts, adaptation, and vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. M. Parry, O. F. Canziani, J. Palutikof, P. J. van der Linden, and C. E. Hanson. Cambridge, UK: Cambridge University Press.

Page 134: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

134

Ehrhart, C., A. Thow, M. de Blois, and A. Warhust. 2008. Humanitarian implications of climate change: Mapping emerging trends and risk hotspots. Chatelaine, Geneva: CARE International.

Erda, L., W. Xiong, H. Ju, Y. L. Xu, Y. Li, L. P. Bai, and L. Y. Xie. 2005. Climate change impacts on crop yield and quality with CO2 fertilization in China. Philosophical Transactions of the Royal Society B-Biological Sciences 360 (1463): 2149–54.

ESCAP (Economic and Social Commission for Asia and the Pacific). 2009. Status of the ESCAP multi-donor voluntary trust fund on tsunami early warning arrangements in the Indian Ocean and Southeast Asia. Background note. http://www.unescap.org/icstd/cdrr/2009/documents/English/CDR_INF6.pdf (accessed April 10, 2009).

estimates. Environmental and Resource Economics (21): 47-53.

Ewing, M., and S. Msangi. 2008. Biofuels production in developing countries: Assessing tradeoffs in welfare and food security. Environmental Science and Policy (forthcoming) doi:10.1016/j.envsci.2008.10.002.

Faisal, I. M., and S. Parveen. 2004. Food security in the face of climate change, population growth, and resource constraints: Implications for Bangladesh. Environmental Management 34 (4): 487–98.

Falcon, W. P., R. L. Naylor, W. L. Smith, M. B. Burke, and E. McCullough. 2004. Using climate models to improve Indonesian food security. Bulletin of Indonesian Economic Studies 40 (3): 355-377

Fan, S., P. Hazell, and S. Thorat. 1999. Linkages between government spending, growth and poverty in rural India. IFPRI Research Report 110. International Food Policy Research Institute, Washington, DC.

Fankhauser, S. 1997. The costs of adapting to climate change. Working Paper No. 13, Global Environmental Facility, Washington, DC.

FAO (Food and Agriculture Organization of the United Nations). 1994. Land degradation in South Asia: Its severity, causes, and effects upon the people. Rome.

FAO (Food and Agriculture Organization of the United Nations). 1994. Land degradation in South Asia: Its severity, causes, and effects upon the people. Rome.

FAO (Food and Agriculture Organization of the United Nations). 2009. Enabling agriculture to contribute to climate change mitigation. An FAO submission to the UNFCCC Ad Hoc Working Group on Long-Term Cooperative Action. Rome.

FAO (Food and Agriculture Organization of the United Nations). 2009. Enabling agriculture to contribute to climate change mitigation. An FAO submission to the UNFCCC Ad Hoc Working Group on Long-Term Cooperative Action. Rome.

FAO (Food and Agriculture Organization of the United Nations). 2001. Integrated agriculture-aquaculture: A primer. FAO Fisheries Technical Paper 407. Rome.

FAOSTAT. 2009. Food and Agriculture Organization of the United Nation Online Statistical Database. http://faostat.fao.org/ (assessed March 20, 2009).

Page 135: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

135

Ferrentino, C. 2007. Potentials for bioenergy to mitigate climate change: A summary of recent findings from the IPCC and the IEA. Rome: Global Bioenergy Partnership Secretariat.

Fischer, G., M. Shah, and H. T. Van Velthuizen. 2002. Climate change and agricultural vulnerability. Laxenburg, Austria: International Institute for Applied Systems Analysis.

Fischer, G., M. Shah, F. Tubiello, and H. T. Van Velthuizen. 2005. Socio-economic and climate change impacts on agriculture: an integrated assessment, 1990-2080. Phil.Trans.R.Soc.B 360: 2067-2083

Fischer, G., M. Shah, F. Tubiello, and H. T. Van Velthuizen. 2005. Socio-economic and climate change impacts on agriculture: An integrated assessment, 1990–2080. Philosophical Transactions of the Royal Society B Biological Sciences 360: 2067–83.

Fischer, G., M. Shah, H. T. Van Velthuizen, and F. O. Nachtergaele. 2001. Global agro-ecological assessment for agriculture in the 21st century. Laxenburg, Austria: International Institute for Applied Systems Analysis.

Fischer, G., M. Shah, H. T. Van Velthuizen, and F. O. Nachtergaele. 2001. Global agro-ecological assessment for agriculture in the 21st Century. Laxenburg, Austria: International Institute for Applied Systems Analysis.

Folke, C., S. Carpenter, T. Elmqvist, L. Gunderson, CS Holling, and B. Walker. 2002. Resilience and Sustainable Development: Building Adaptive Capacity in a World of Transformations. AMBIO 31 (5): 437-440

Folke, C., S. Carpenter, T. Elmqvist, L. Gunderson, CS Holling, and B. Walker. 2002. Resilience and Sustainable Development: Building Adaptive Capacity in a World of Transformations. AMBIO 31 (5): 437-440

Francisco, H. A. 2008. Adaptation to climate change: Needs and opportunities in Southeast Asia. ASEAN Economic Bulletin 25 (1): 7–19.

Fussel, H. M. 2007. Vulnerability: A generally applicable conceptual framework for climate change research. Global Environmental Change 17 (2): 155–67.

Gbetibouo, G.A. 2009. Mapping South African Farming Sector Vulnerability to Climate Change and Variability. IFPRI Discussion Paper, under review.

GEF (Global Environmental Facility). 2008. The Least Developed Countries Fund. Washington, DC.

Gillett, R., M. McCoy, L. Rodwell, and J. Tamate. 2001. Tuna: A key economic resource in the Pacific. Pacific Studies Series. Manila: Asian Development Bank.

Giordano, M., and K. Villholth. 2007. The agricultural groundwater revolution: Setting the stage. In The agricultural groundwater revolution: Opportunities and threats to development, ed. M. Giordano and K. Villholth. Wallingford, UK: CAB International.

Government of the Republic of Korea. 2003. Second national communication of the Republic of Korea under the United Nations Framework Convention on Climate Change. Seoul.

Gruhn, P., F. Goletti, and M. Yudelman. 2000. Integrated nutrient management, soil fertility, and sustainable agriculture: Current issues and future challenges. 2020 Discussion Paper 32. Washington, DC: International Food Policy Research Institute.

Page 136: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

136

Hakimov, A., A. Lines, P. Elmuratov, and R. Hakimov. 2007. Climate change and water resource alteration in Central Asia: The case of Uzbekistan. In Climate change and terrestrial carbon sequestration in Central Asia, ed. R. Lal, M. Suleimenov, B. A. Steward, D. O. Hansen, and P. Doraiswamy. Oxford: Routledge.

Hales, S., N. de Wet, J. Maindonald, and A. Woodward. 2002. Potential effect of population and climate changes on global distribution of dengue fever: An empirical model. Lancet 360 (9336): 830–34.

Hales, S., P. Weinstein, Y. Souares, and A. Woodward. 1999. El Niño and the dynamics of vectorborne disease transmission. Environmental Health Perspectives 107 (2): 99–102.

Hales, S., S. J. Edwards, and R. S. Kovats. 2003. Impacts on health of climate extremes. In Climate change and human health: Risks and responses, ed. A. J. McMichael, D. H. Campbell-Lendrum, and C. F. Corvalanet. Geneva: World Health Organization.

Harmeling, S. 2008. Global climate risk index 2009: Weather-related loss eventsand their impacts on countries in 2007 and in a long-term comparison. Bonn: Germanwatch.

Herath, T. N. 2001. Farmer's knowledge of water management methods in the dry zone of Sri Lanka. Indigenous Knowledge and Development Monitor 9 (3): 27–28. http://web.archive.org/web/20040719195342/www.nuffic.nl/pdf/ciran/ikdm0111.pdf.

Hijmans, R.J., S.E. Cameron, J.L. Parra, P.G. Jones and A. Jarvis, 2005. Very high resolution interpolated climate surfaces for global land areas. International Journal of Climatology 25: 1965-1978

Hoerling, M., and A. Kumar. 2003. The perfect ocean for drought. Science 299 (5607): 691-694

Holling, C. S. 2001. Understanding the complexity of economic, ecological, and social systems. Ecosystems 4:390–405.

Hopp, M. J., and J. A. Foley. 2003. Worldwide fluctuations in dengue fever cases related to climate variability. Climate Research 25 (1): 85–94.

Howden, S. M., J. Soussana, F. Tubiello, N. Chhetri, and M. Dunlop. 2007. Adapting agriculture to climate change. PNAS 104 (50): 19691-19696

Hughes, B.B., M.T. Irfan, H. Khan, K.B. Kumar, D.S. Rothman, J.R. Solorzano. 2008. Reducing global poverty: Patterns of potential human progress. Volume 1. The Pardee Center for International Futures, Oxford University Press.

ICRISAT (International Crops Research Institute for the Semi-Arid Tropics). 2007. Pro-poor biofuels outlook for Asia and Africa. Working Paper. Patancheru, India.

IEA (International Energy Agency). 2006. World energy outlook. Paris.

IGES (Institute for Global Environmental Strategies). 2008. Climate change policies in the Asia-Pacific: Re-uniting climate change and sustainable development. http://www.iges.or.jp.

IPCC (Intergovernmental Panel on Climate Change). 2007. Annex 1: Glossary. In Fourth assessment report, Working Group 2. http://www.ipcc.ch/pdf/glossary/ar4-wg2.pdf (accessed March 29, 2009).

IPCC. 2007. Summary for Policymakers. In Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the

Page 137: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

137

Intergovernmental Panel on Climate Change. Ed. M. Parry, O. F. Canziani, J. Palutikof, P. J. van der Linden, and C. E. Hanson. Cambridge: Cambridge University Press.

Jamting, H. 2008. Gender and climate change. Uppsala. Swedish University of Agricultural Science.

Ji, C. Y. 2008. Land degradation in Central Asia. ADB TA 6356-REG: Central Asian Countries Initiative for Land Management Multicountry Partnership Framework Support Project. Asian Development Bank, Manila.

Jose, A., and N. A. Cruz. 1999. Climate change impacts and responses in the Philippines: Water resources. Climate Research 12 (2–3): 77–84.

Kahn, M. E. 2005. The death toll from natural disasters: The role of income, geography, and institutions. Review of Economics and Statistics 87 (2): 271-284

Kahn, M. E. 2005. The death toll from natural disasters: The role of income, geography, and institutions. Review of Economics and Statistics 87 (2): 271–84.

Kang Su, K., and L. Chung Kuen. 2006. How will the increased air temperature by global warming influence on the yield capacity and quality of rice in Korea? Nippon Sakumotsu Gakkai Koenkai Yoshi, Shiryoshu 221: 414–15.

Karim, Z., G. Hussain, and A. U. Ahmed. 1999. Climate change vulnerability of crop agriculture. In Decision criteria and optimal inventory processes, ed. B. Liu and A. O. Esogbue. New York, Springer.

Karim, Z., S. G. Hussain, and M. Ahmed. 1996. Assessing impacts of climatic variations on foodgrain production in Bangladesh. Water Air and Soil Pollution 92 (1–2): 53–62.

Kelkar, U., and S. Bhadwal. 2007. South Asian Regional Study on Climate Change Impacts and Adaptation: Implications for Human Development. Occasional Paper No. 2007/27. UNDP

Kelkar, U., and S. Bhadwal. 2007. South Asian regional study on climate change impacts and adaptation: Implications for human development. Occasional Paper No. 2007/27. New York: United Nations Development Programme.

Khan, H. U., and T. Kurosaki. 2007. Vulnerability of microfinance to natural disasters: Evidence from the 2005 Pakistan earthquake. Draft.

Khosla, P., and R. Pearl. 2003. Gender, water, and poverty: Key issues, government commitments and actions for sustainable development. New York: WEDO -Women's Environment and Development Organization

Kjellstrom, T., R. S. Kovats, S. J. Lloyd, T. Holt, and R. S. J. Tol. 2008. The direct impact of climate change on regional labour productivity. Working Paper No. 260. Dublin: ESRI. Economic and Social Research Institute

Kumar, K., and J. Parikh. 1998. Climate change impacts on Indian agriculture: The Ricardian approach. In Measuring the impact of climate change on Indian agriculture. World Bank Technical Paper. Washington, DC: World Bank.

Kundzewicz, Z. W., L. J. Mata, N. W. Arnell, P. Doll, P. Kabat, B. Jimenez, K. A. Miller, T. Oki, Z. Sen, and I. A. Shiklomanov. 2007. Freshwater resources and their management.

Page 138: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

138

In Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Ed. M. Parry, O. F. Canziani, J. Palutikof, P. J. van der Linden, and C. E. Hanson. Cambridge: Cambridge University Press.

Kundzewicz, Z. W., L. J. Mata, N. W. Arnell, P. Doll, P. Kabat, B. Jimenez, K. A. Miller, T. Oki, Z. Sen, and I. A. Shiklomanov. 2007. Freshwater resources and their management. In Climate change 2007: Impacts, adaptation, and vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. M. Parry, O. F. Canziani, J. Palutikof, P. J. van der Linden, and C. E. Hanson. Cambridge, UK: Cambridge University Press.

Kurukulasuriya, P., and M. I. Ajwad. 2007. Application of the Ricardian technique to estimate the impact of climate change on smallholder farming in Sri Lanka. Climatic Change 81 (1): 39-59

Kurukulasuriya, P., and M. I. Ajwad. 2007. Application of the Ricardian technique to estimate the impact of climate change on smallholder farming in Sri Lanka. Climatic Change 81 (1): 39–59.

Lal, R, and J. P. Bruce. 1999. The potential of world cropland soils to sequester C and mitigate the greenhouse effect. Environmental Science and Policy 2, 177–85.

Lal, R. 2004. The potential of carbon sequestration in the soils of South Asia. Paper No. 134. Presented at the 13th International Soil Conservation Organization Conference, Brisbane, Australia, July.

Lambrou, Y., and G. Piana. 2006. Gender: The missing component of the response to climate change. Rome: Food and Agriculture Organization of the United Nations.

Lansigan, F. 2003. Assessing the impacts of climate variability on crop production and developing coping strategies in rainfed agriculture. In Coping against El Nino for stabilizing rainfed agriculture: Lessons from Asia and the Pacific, ed. S. Yokoyama and R. N. Concepcion. Proceedings of a joint workshop, Cebu, Philippines, September 17–19, 2002. Bangkok: United Nations Economic and Social Commission for Asia and the Pacific (ESCAP).

Leary, N. A., J.Adejuwon, V. Barros, P. Batimaa, B. Biagini, I. Burton, S. Chinvanno, R. Cruz, D. Dabi, A. de Comarmond, B. Dougherty, P. Dube, A. Githeko, A. Abou Hadid, M. Hellmuth, R. Kangalawe, J. Kulkarni, M. Kumar, R. Lasco, M. Mataki, M. Medany, M. Mohsen, G. Nagy, M. Njie, J.Nkomo, A. Nyong, B. Osman, E. Sanjak, R. Seiler, M. Taylor, M. Travasso, G. von Maltitz, S. Wandiga and M. Wehbe. 2007. A Stitch in Time: Lessons for Climate Change Adaptation from the AIACC Project, AIACC Working Paper 48, Washington, D.C. [Available online at http://www.aiaccproject.org].

Leary, N.A. 1999. A framework for benefit-cost analysis of adaptation to climate change and climate variability. Mitigation and Adaptation Strategies for Global Change 4 (3–4), 307–318.

Lobell, D. B., M. B. Burke, C. Tebaldi, M. D. Mastrandrea, W. P. Falcon, and R. L. Naylor. 2008. Prioritizing climate change adaptation needs for food security in 2030. Science 319 (5863): 607-610

Page 139: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

139

Lobell, D. B., M. B. Burke, C. Tebaldi, M. D. Mastrandrea, W. P. Falcon, and R. L. Naylor. 2008. Prioritizing climate change adaptation needs for food security in 2030. Science 319 (5863): 607–10.

MA (Millennium Ecosystem Assessment). 2005. Ecosystems and human well-being: Scenarios. Millennium Ecosystem Assessment Series, Vol. 2. Washington, DC: Island Press.

MA (Millennium Ecosystem Assessment). 2005. Ecosystems and human well-being: Scenarios. Millennium Ecosystem Assessment Series, Vol. 2. Washington, DC: Island Press.

Magrath, J., S. Carvalho, R. M. Nkuba, A. Nkubito, and J. Wangoolo. 2008. Turning up the heat: Climate change and poverty in Uganda. Uganda: Oxfam.

Mallick, F. 2006. Adaptation in action: Community level adaptation techniques in the context of the southwestern region of Bangladesh - The Reducing Vulnerability to Climate Change (RVCC) Project, Presentation at the International Workshop on Community Level Adaptation to Climate Change, Dhaka: Care Bangladesh

Manandhar, N. P. 1998. The preparation of gundruk in Nepal: A sustainable rural industry? Indigenous Knowledge and Developmetn Monitor 6 (3). http://web.archive.org/web/20041215133649/www.nuffic.nl/ciran/ikdm/6-3/manandh.html.

Manton, M. J., P. M. la-Marta, M. R. Haylock, K. J. Hennessy, N. Nicholls, L. E. Chambers, D. A. Collins, G. Daw, A. Finet, D. Gunawan, K. Inape, H. Isobe, T. S. Kestin, P. Lefale, C. H. Leyu, T. Lwin, L. Maitrepierre, N. Ouprasitwong, C. M. Page, J. Pahalad, N. Plummer, M. J. Salinger, R. Suppiah, V. L. Tran, B. Trewin, I. Tibig, and D. Yee. 2001. Trends in extreme daily rainfall and temperature in Southeast Asia and the South Pacific: 1961-1998. International Journal of Climatology 21 (3): 269-284

Matthews, R. B., M. J. Kropff, T. Horie, and D. Bachelet. 1997. Simulating the impact of climate change on rice production in Asia and evaluating options for adaptation. Agricultural Systems 54 (3): 399-425

McCarthy, J., O. F. Canziani, N. A. Leary, D. J. Dokken, and K. S. White, ed. 2001. Climate change 2001: Impacts, adaptation, and vulnerability. Contribution of Working Group II to the Third Assessment . Cambridge: Cambridge University Press

McMichael, A. J., R. E. Woodruff, and S. Hales. 2006. Climate change and human health: Present and future risks. Lancet 367 (9513): 859–69.

Meehl, G. A., T. F. Stocker, W. D. Collins, P. Friedlingstein, A. T. Gaye, J. M. Gregory, A. Kitoh, R. Knutti, J. M. Murphy, A. Noda, S. C. B. Raper, I. G. Watterson, A. J. Weaver, and Z. Zhao. 2007. Global Climate Projections. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Ed. S. Solomon, Qin.D., M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, and H. L. Miller. Cambridge: Cambridge University Press.

Meinzen-Dick, R., and K. Moore. 2009. Institutional dimensions of climate change: The role of property rights and collective action. CAPRi Working Paper. Washington, DC: International Food Policy Research Institute.

Page 140: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

140

Mendelsohn, R. 2005. Climate change impacts on Southeast Asian agriculture. Yale University.

Mendelsohn, R. 2006. The role of markets and governments in helping society adapt to a changing climate. Climatic Change 78 (1): 203-215

Mendelsohn, R., and L. Williams. 2004. Comparing forecasts of the global impacts of climate change. Mitigation and Adaptation Strategies for Global Change 9: 315-333

Mendelsohn, R., and L. Williams. 2004. Comparing forecasts of the global impacts of climate change. Mitigation and Adaptation Strategies for Global Change 9 (4): 315–33.

Mendelsohn, R., W. Morrison, M. E. Schlesinger, and N. G. Andronova. 2000. Country-specific market impacts of climate change. Climatic Change 45 (3–4): 553–69.

Meusch, E., J. Yhoung-Aree, R. Friend, and S. J. Funge-Smith. 2003. The role and nutritional value of aquatic resources in the livelihoods of rural people: A participatory assessment in Attapeu Province, Lao PDR. Publication No. 2003/11. Bangkok: Food and Agriculture Organization of the United Nations, Regional Office for Asia and the Pacific.

Mirza, M. 2007. Climate Change, Adaptation and Adaptive Governance in Water Sector in South Asia. Toronto: Environment Canada

Mirza, M. 2007. Climate change, adaptation, and adaptive governance in water sector in South Asia. Toronto: Environment Canada.

Mitchell, T., T. Tanner, and K. Lussier. 2007. We know what we need: South Asian women speak out on climate change adaptation. London: Actionaid and Institute of Development Studies.

Mitchell, T.D. and P.D., Jones (2005), An improved method of constructing a database of monthly climate observations and associated high-resolution grids, International Journal of Climatology, 25: 693-712.

Mitchell, T.D., T.R., Carter, P.D., Jones, M., Hulme and M., New (2004), A comprehensive set of high-resolution grids of monthly climate for Europe and the globe: the observed record (1901-2000) and 16 scenarios (2001-2100), Tyndall Center Working paper No. 55.

Mosier, A., C. Kroeze, C. Nevison, O. Oenema, S. Seitzinger and O. van Cleemput. 1998. Closing the global N2O budget: Nitrous oxide emissions through the agricultural nitrogen cycle. Nutrient Cycling in Agroecosystems 52 (2–3): 225–48.

Moss, R., A. Brenkert, and E. Malone. 2001. Vulnerability to climate change, A quantitative Approach. NTIS

Moss, R., A. Brenkert, and E. Malone. 2001. Vulnerability to climate change: A quantitative Approach. Springfield, Virginia: The National Technical Information Service

MPND (Ministry of Planning and National Development). 2007. Seventh national development plan, 2006–2010. Male: MPND, Republic of Maldives.

Murray, B. C., B. Sohngen, A. J. Sommer, K. Jones, B. McCarl, D. Gillig, B. DeAngelo, and K. Andrasko. Andrasko, K., and DeAngelo, B., ed. 2005. Greenhouse Gas Mitigation Potential in U.S. Forestry and Agriculture. Washington, DC: US EPA

Page 141: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

141

Narain, P., M. A. Khan, and G. Singh. 2005. Potential for water conservation and harvesting against drought in Rajasthan, India. Working Paper 104 (Drought Series: Paper 7). Colombo, Sri Lanka: International Water Management Institute (IWMI).

Naylor, R. L., D. S. Battisti, W. P. Falcon, and M. B. Burke. 2007. Assessing risks of climate variability and climate change for Indonesian rice agriculture. PNAS 104 (19): 7752-7757

Neumayer, E., and T. Plumper. 2007. The gendered nature of natural disasters: The impact of catastrophic events on the gender gap in life expectancy, 1981–2002. Annals of the Association of American Geographers 97 (3): 551–66.

Nguyen, H. N. 2007. Flooding in Mekong River Delta, Viet Nam. Occasional Paper No. 2007/53. New York: United Nations Development Programme.

Nguyen, N. V. 2005. Global climate changes and rice food security. Rome: Food and Agriculture Organization of the United Nations.

Nguyen, N. V. 2005. Global climate changes and rice food security. Rome: Food and Agriculture Organization of the United Nations.

Noy, I. 2009. The macroeconomic consequences of disasters. Journal of Development Economics 88 (2): 221-231

Noy, I. 2009. The macroeconomic consequences of disasters. Journal of Development Economics 88 (2): 221–31.

Noy, I. 2009. The macroeconomic consequences of disasters. Journal of Development Economics 88 (2): 221–31.

O'Brien, K. 2000. Developing strategies for climate change: The UNEP country studies on climate change impacts and adaptations assessment. CICERO Report No. 2000:2. Oslo: Center for International Climate and Environmental Research (CICERO) and United Nations Environment Programme.

O'Brien, K. 2000. Developing strategies for climate change: The UNEP country studies on climate change impacts and adaptations assessment. CICERO Report No. 2000:2. Oslo: Center for International Climate and Environmental Research (CICERO) and United Nations Environment Programme.

O'Brien, K. L., S. Eriksen, L. Nygaard, and A. Schjolden. 2007. Different interpretations of vulnerability matter in climate change discourses. Climate Policy 7 (1): 73–88.

O'Brien, K., R. Leichenko, U. Kelkar, H. Venema, G. Aandahl, H. Tompkins, A. Javed, S. Bhadwal, S. Barg, L. Nygaard, and J. West. 2004. Mapping vulnerability to multiple stressors: Climate change and globalization in India. Global Environmental Change Part A 14 (4): 303–13.

Oenema, O., N. Wrage, G. L. Velthof, J. W. van Groenigen, J. Dolfing and P. J. Kuikman. 2005. Trends in global nitrous oxide emissions from animal production systems. Nutrient Cycling in Agroecosystems 72 (1): 51–65.

Oxfam and Kyoto University. 2007. Drought-management considerations for climate change adaptation: Focus on the Mekong region. Kyoto: Kyoto University.

Page 142: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

142

Oxfam. 2008. Viet Nam: Climate Change, Adaptation and Poor People. Viet Nam: Oxfam

Pacala, S., and R. Socolow. 2004. Stabilization wedges: Solving the climate problem for the next 50 years with current technologies. Science 305 (5686): 968–72.

Pahl-Wostl, C. 2007a. The implications of complexity for integrated resources management. Environmental Modelling and Software 22 (5): 561-569

Pahl-Wostl, C. 2007a. The implications of complexity for integrated resources management. Environmental Modelling and Software 22 (5): 561-569

Pahl-Wostl, C. 2007b. Transitions towards adaptive management of water facing climate and global change. Water Resources Management 21 (1): 49-62

Pahl-Wostl, C. 2007b. Transitions towards adaptive management of water facing climate and global change. Water Resources Management 21 (1): 49-62

Pandey, S., H. Bhandari, S. Ding, P. Prapertchob, R. Sharan, D. Naik, S. K. Taunk, and A. Sastri. 2007. Coping with drought in rice farming in Asia: Insights from a cross-country comparative study. Agricultural Economics 37 (S1): 213–24.

Pandya-Lorch, R., and M. W. Rosegrant. 2000. Prospects for food demand and supply in Central Asia. Food Policy 25 (6): 637–46.

Parikh, J. 2007. Gender and climate change: Framework for analysis, policy, and action. New Delhi, India: United Nations Development Programme and Integrated Research and Action for Development (IRADe).

Parry, M. 2007. The implications of climate change for crop yields, global food supply, and risk of hunger. ICRISAT 4 (1) 44p.

Parry, M., C. Rosenzweig, A. Iglesias, G. Fischer, and M. Livermore. 1999. Climate change and world food security: A new assessment. Global Environmental Change 9 (S1): S51–S67.

Parry, M., C. Rosenzweig, A. Iglesias, M. Livermore, and G. Fischer. 2004. Effects of climate change on global food and productions under SRES emissions and socio-economic scenarios. Global Environmental Change 14 (1): 53–67.

Parry, M., C. Rosenzweig, A. Iglesias, M. Livermore, and G. Fischer. 2004. Effects of climate change on global food and productions under SRES emissions and socio-economic scenarios. Global Environmental Change 14 (1): 53–67.

Parry, M., C. Rosenzweig, and M. Livermore. 2005. Climate change, global food supply and risk of hunger. Philosophical Transactions of the Royal Society 360 (1463): 2125–38.

Parry, M., C. Rosenzweig, and M. Livermore. 2005. Climate change, global food supply, and risk of hunger. Philosophical Transactions of the Royal Society 360 (1463): 2125–38.

Pathak, H., and R. Wassmann. 2007. Introducing greenhouse gas mitigation as a development objective in rice-based agriculture: I. Generation of technical coefficients. Agricultural Systems 94 (3): 807–25.

Patt, A., R. J. T. Klein, and A. Vega-Leinert. 2005. Taking the uncertainty in climate-change vulnerability assessment seriously. Comptes Rendus Geoscience 337 (4): 411–24.

Page 143: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

143

Patz, J. A., D. Campbell-Lendrum, T. Holloway, and J. A. Foley. 2005. Impact of regional climate change on human health. Nature 438 (7066): 310–17.

Peng, S., J. Huang, J. Sheehy, R. Laza, R. Visperas, X. Zhong, G. Centeno, G. Khush, and K. Cassman. 2004. Rice yields decline with higher night temperature from global warming. PNAS (Proceedings of the National Academy of Sciences) 101 (27): 9971–75.

Peralta, A. 2008. Gender and climate change finance: A case study from the Philippines. New York: Women's Environment and Development Organization

Perelet, R. 2007. Central Asia: Background paper on climate change. Occasional Paper 2007/11. New York: United Nations Development Programme.

Pimentel, D., and T. Patzek. 2005. Ethanol production using corn, switchgrass, and wood; biodiesel production using soybean and sunflower. Natural Resources Research 14 (1): 65–76.

Pimentel, D., P. Hepperly, J. Hanson, D. Douds, and R. Seidel. 2005. Environmental, energetic, and economic comparisons of organic and conventional farming systems. BioScience 55 (7): 573–82.

Pittock, B. and R.N. Jones. 2000. Adaptation to what and why? Environmental Monitoring and Assessment 61(1), 9–35.

Poverty-Environment Partnership. 2003. Poverty and Climate Change. Reducing the Vulnerability of the Poor through Adaptation. UNDP, UNEP, World Bank, ADB, AfDB, GTZ, DFID, OECD, EC

Poverty-Environment Partnership. 2003. Poverty and climate change: Reducing the vulnerability of the poor through adaptation. African Development Bank; Asian Development Bank; Department for International Development, United Kingdom; Directorate-General for Development, European Commission; Federal Ministry for Economic Cooperation and Development, Germany; Ministry of Foreign Affairs - Development Cooperation, The Netherlands; Organization for Economic Cooperation and Development; United Nations Development Programme; United Nations Environment Programme; and World Bank.

Poverty-Environment Partnership. 2003. Poverty and climate change: Reducing the vulnerability of the poor through adaptation. African Development Bank; Asian Development Bank; Department for International Development, United Kingdom; Directorate-General for Development, European Commission; Federal Ministry for Economic Cooperation and Development, Germany; Ministry of Foreign Affairs - Development Cooperation, The Netherlands; Organization for Economic Cooperation and Development; United Nations Development Programme; United Nations Environment Programme; and World Bank.

Preston, B. L., R. Suppiah, I. Macadam, and J. Bathols. 2006. Climate Change in the Asia/Pacific Region. A consultancy report prepared for the climate change and development roundtable. Australia: CSIRO

Preston, B. L., R. Suppiah, I. Macadam, and J. Bathols. 2006. Climate change in the Asia/Pacific region: A consultancy report prepared for the climate change and development roundtable. Clayton, Australia: Commonwealth Scientific and Industrial Research Organisation.

Page 144: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

144

Preston, B. L., R. Suppiah, I. Macadam, and J. Bathols. 2006. Climate change in the Asia/Pacific region: A consultancy report prepared for the climate change and development roundtable. Clayton, Australia: Commonwealth Scientific and Industrial Research Organisation.

Quan, J., and N. Dyer. 2008. Climate change and land tenure: The implications of climate change for land tenure and land policy. Land Tenure Working Paper 2. London: International Institute for Environment and Development and the United Nations Food and Agriculture Organization.

RA (Resilience Alliance). 2009. Glossary. http://www.resalliance.org/608.php#S (accessed April 8, 2009).

Rajagopal, D., and D. Zilberman. 2007. Review of environmental, economic, and policy aspects of biofuels. Policy Research Working Paper No. 4341. Washington, DC: World Bank.

Rao, C. H. 1994. Agricultural growth, rural poverty, and environmental degradation in India. Delhi: Oxford University Press.

Ray, I. 2007. Women, water, and development. Annual Review of Environment and Resources 32 (1): 421–49.

Reiter, P. 2001. Climate change and mosquito-borne disease. Environmental Health Perspectives 109 (S1): 141–61.

Renault, D., T. Facon, and R. Wahaj. 2007. Modernizing irrigation management—The MASSCOTE approach: Mapping system and services for canal operation techniques. FAO Irrigation and Drainage Paper No. 63. Rome: Food and Agriculture Organization of the United Nations.

Reske-Nielsen, F. 2008. Climate change: Timor-Leste confronting the environmental challenges of the developing world. News Quarterly (United Nations Development Programme–Timor Leste) 11 (June): 2.

Reti, M. J. 2008. An assessment of the impact of climate change on agriculture and food security in the Pacific: A case study in the Republic of the Marshall Islands. Samoa: Pacific Environment Consultants Ltd.

Robert, M. 2001. Soil carbon sequestration for improved land management. Rome: Food and Agriculture Organization of the United Nations.

Rogers, D. J., and S. E. Randolph. 2000. The global spread of malaria in a future, warmer world. Science 289 (5485): 1763–66.

Rosegrant, M. W. 2008. Biofuels and grain prices: Impacts and policy responses. Testimony for the U.S. Senate Committee on Homeland Security and Governmental Affairs. Washington, DC: International Food Policy Research Institute.

Rosegrant, M.W., M.S.Paisner, S. Meijer and J. Witcover. 2001. Global Food Projections to 2020: Emerging Trends and Alternative Futures. 2020 Vision Food Policy Report. Washington D.C., International Food Policy Research Institute.

Page 145: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

145

Rosenzweig, C., A. Iglesias, X. Yang, P. R. Epstein, and E. Chivian. 2001. Climate change and extreme weather events: Implications for food production, plant diseases, and pests. Global Change and Human Health 2 (2): 90–104.

Rossi, A., and Y. Lambrou. 2008. Gender and equity issues in liquid biofuels production: Minimizing the risks to maximize the opportunities. Rome: Food and Agriculture Organization of the United Nations.

Sachs, C. 2009. Module 10: Gender and natural resources management. In Gender in agriculture sourcebook. Washington, DC: World Bank, Food and Agriculture Organization of the United Nations, and International Fund for Agricultural Development.

Sachs, J., and P. Malaney. 2002. The economic and social burden of malaria. Nature 415 (6872): 680–85.

Sanghi, A., R. Mendelsohn, and A. Dinar. 1998. The climate sensitivity of Indian agriculture. In Measuring the impact of climate change on Indian agriculture, ed. A. Dinar, R. Mendelsohn, R. Evenson, J. Parikh, A. Sanghi, K. Kumar, J. McKinsey, and S. Lonergan. Technical Paper 402. Washington, DC: World Bank.

Sanker, S., H. Nakano, and Y. Shiomi. 2007. Natural Disasters Data Book - 2006. An Analytical Overview . Kobe, Japan: ADRC

Sanker, S., H. Nakano, and Y. Shiomi. 2007. Natural disasters data book— 2006: An analytical overview . Kobe, Japan: Asian Disaster Reduction Center.

Sanker, S., H. Nakano, and Y. Shiomi. 2007. Natural disasters data book—2006: An analytical overview . Kobe, Japan: Asian Disaster Reduction Center.

Sari, A. P., M. Maulidya, R. N. Butarbuta, R. E. Sari, and W. Rusmantoro. 2007. Executive summary: Indonesia and climate change: Working paper on current status and policies. Jakarta: Pelangi Energi Abadi Citra Enviro - PEACE.

Saseendran, S. A., K. K. Singh, L. S. Rathore, S. V. Singh, and S. K. Sinha. 2000. Effects of climate change on rice production in the tropical humid climate of Kerala, India. Climatic Change 44 (4): 495–514.

Schmidhuber, J., and F. Tubiello. 2007. Global food security under climate change. PNAS (Proceedings of the National Academy of Sciences) 104 (50): 19703–08.

Schubert, R., H. J. Schellnhuber, N. Buchmann, Griebhamme R., M. Kulessa, D. Messner, D. Messner, S. Rahmstorf, and J. Schmid. 2008. Climate change as a security risk. London: German Advisory Council on Global Change and Earthscan.

Schubert, R., H. J. Schellnhuber, N. Buchmann, Griebhamme R., M. Kulessa, D. Messner, D. Messner, S. Rahmstorf, and J. Schmid. 2008. Climate change as a security risk. London: German Advisory Council on Global Change and Earthscan

Searchinger, T., R. Heimlich, R.A. Houghton, F. Dong, A. Elobeid, J. Fabiosa, S. Tokgoz, D. Hayes, and T. Yu. 2008. Use of U.S. croplands for biofuels increases greenhouse gases through emissions from land-use change. Science 319 (5687): 1238.

Page 146: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

146

Selvaraju, A., R. Subbiah, S. Baas, and I. Juergens. 2006. Livelihood adaptation to climate variability and change in drought-prone areas in Bangladesh. Rome: Asian Preparedness Centre and Food and Agriculture Organization of the United Nations.

Seo, S. N., R. Mendelsohn, and M. Munasinghe. 2005. Climate change and agriculture in Sri Lanka: A Ricardian valuation. Environment and Development Economics 10 (5): 581–96.

Shagdar, E. 2002. The Mongolian livestock sector: Vital for the economy and people, but vulnerable to natural phenomena. Niigata, Japan: Economic Research Institute for Northeast Asia.

Sharan, D. 2008. Financing climate change mitigation and adaptation: Role of regional financing arrangements. Sustainable Development Working Paper 4. Manila: Asian Development Bank.

Sharan, D. 2008. Financing climate change mitigation and adaptation: Role of regional financing arrangements. Sustainable Development Working Paper 4. Manila: Asian Development Bank.

Shea, E. 2001. Challenges and opportunities. In Preparing for a changing climate: The potential consequences of climate variability and change. Pacific Islands. Honolulu, HI, USA: East-West Center.

Shrestha, S., and K. Yatsuka. 2008. Groundwater and climate change: No longer the hidden resource. In Climate change policies in the Asia-Pacific: Re-uniting climate change and sustainable development. Hayama, Japan: Institute for Global Environmental Strategies.

Singh, R. B. K., S. Hales, N. de Wet, R. Raj, M. Hearnden, and P. Weinstein. 2001. The influence of climate variation and change on diarrheal disease in the Pacific Islands. Environmental Health Perspectives 109 (2): 155–59.

Sivanappan, R. K. 1997. Technologies for water harvesting and soil moisture conservation in small watersheds for small-scale irrigation. Rome: Food and Agriculture Organization of the United Nations.

Smit, B. and O. Pilifosova (2001) Adaptation to climate change in the context of sustainable development and equity (Chapter 18). In: McCarthy, J.J., O.F. Canziani, N.A. Leary, D.J. Dokken, and K.S. White (eds.) (2001) Climate Change 2001: impacts, adaptation and vulnerability. Contributions of Working Group II, Cambridge University Press, Cambridge, UK.

Smith, K. A., and F. Conen. 2004. Impacts of land management on fluxes of trace greenhouse gases. Soil Use and Management 20 (2): 255–63.Trends in Global Nitrous Oxide Emissions from Animal Production Systems

Smith, L. and L. Haddad. 2000. Explaining Child Malnutrition in Developing Countries: A Cross-Country Analysis. IFPRI Research Report. IFPRI: Washington, DC.

Smith, P. 2009. Costs of mitigation in sub-regions of Asia at various carbon prices. Unpublished data. University of Aberdeen.

Smith, P., D. Martino, Z. Cai, D. Gwary, H. H. Janzen, P. Kumar, B. McCarl, S. Ogle, F. O’Mara, C. Rice, R. J. Scholes, O. Sirotenko, M. Howden, T. McAllister, G. Pan, V.

Page 147: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

147

Romanenkov, U. Schneider, S. Towprayoon, M. Wattenbach, and J. U. Smith. 2007b. Greenhouse gas mitigation in agriculture. Philosophical Transactions of the Royal Society B Biological Sciences 363 (1492): 789–813.

Smith, P., D. Martino, Z. Cai, D. Gwary, H. Janzen, P. Kumar, B. McCarl, S. Ogle, F. O’Mara, C. Rice, B. Scholes, and O. Sirotenko. 2007a. Agriculture. In Climate change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. B. Metz, O. R. Davidson, P. R. Bosch, R. Dave, and L. A. Meyer. Cambridge, UK: Cambridge University Press.

Snidvongs, A. 2006. Vulnerability to climate change related water resource changes and extreme hydrological events in Southeast Asia. A final report submitted to Assessments of Impacts and Adaptations to Climate Change (AIACC), Project No. AS 07. Washington, DC: International START Secretariat.

Snidvongs, A. 2006. Vulnerability to climate change related water resource changes and extreme hydrological events in Southeast Asia. A final report submitted to Assessments of Impacts and Adaptations to Climate Change (AIACC), Project No. AS 07. Washington, DC: International START Secretariat.

SRI (Stockholm Resilience Institute). 2009. Resilience dictionary. http://www.stockholmresilience.org/research/whatisresilience/resiliencedictionary.4.aeea46911a3127427980004355.html (accessed March 17, 2009).

Steinfeld, H., P. Gerber, T. Wassenaar, V. Castel, M. Rosales and C. De Haan. 2006. Livestock’s long shadow: Environmental issues and options. Rome: Food and Agriculture Organization of the United Nations.

Steinfeld, H., P. Gerber, T. Wassenar, V. Castelo, M. Rosales, and C. De Haan. 2006. Livestock's long shadow: Environmental issues and options. Rome: Food and Agriculture Organization of the United Nations.

Stromberg, D. 2007. Natural disasters, economic development, and humanitarian aid. Journal of Economic Perspectives 21 (3): 199-222

Stromberg, D. 2007. Natural disasters, economic development, and humanitarian aid. Journal of Economic Perspectives 21 (3): 199–222.

Tay, S. 2008. Climate change and insecurity in the Asia Pacific: Regional initiatives to address emerging threats. In Security through cooperation: Furthering Asia Pacific multilateral engagement, ed. B. L. Job. Vancouver: Council for Security Cooperation in the Asia Pacific.

TERI (The Energy and Resources Institute). N.d. Three case studies using different ecosystems. INTEREST (Interactions between Environment, Society and Technology), TERI, New Delhi. http://www.teriin.org/teri-wr/projects/interestaqua.pdf.

Tol, R.S.J. (2002). Estimates of the damage costs of climate change. Part I. Benchmark

Tri, N. H., W. N. Adger, and P. M. Kelly. 1998. Natural resource management in mitigating climate impacts: Mangrove restoration in Viet Nam. Global Environmental Change 8 (1): 49–61.

Page 148: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

148

Tschakert, P. 2007. Views from the vulnerable: Understanding climatic and other stressors in the Sahel. Global Environmental Change 17 (3–4): 381–96.

Tubiello, F., and C. Rosenzweig. 2008. Developing climate change impact metrics for agriculture. Integrated Assessment Journal 8 (1): 165–84.

Tubiello, F., J. Schmidhuber, M. Howden, and E. Fernandes. 2008. Climate Change Response Strategies for Agriculture: Challenges and Opportunities for the 21st Century. Agriculture and Rural Development (ARD) Discussion PaperWashington, DC: World Bank

UNEP DTIE (United Nations Environment Programme, Division of Technology, Industry, and Economics). 2000. Rooftop rainwater harvesting for domestic water supply. In Sourcebook of alternative technologies for freshwater augmentation in some countries in Asia. http://www.unep.or.jp/ietc/publications/techpublications/techpub-8e/rooftop.asp.

UNFCCC (United Nations Framework Convention on Climate Change). 2008a. Increased growing season in central Tajikistan through cold frames. CARE Tajikistan. UNFCCC database on local coping strategies. http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=177.

USEPA (U.S. Environmental Protection Agency). 2006. Global mitigation of non-CO2 greenhouse gases. Washington, DC: Office of Atmospheric Programs, USEPA.

van Lieshout, M., R. S. Kovats, M. T. J. Livermore, and P. Martens. 2004. Climate change and malaria: Analysis of the SRES climate and socio-economic scenarios. Global Environmental Change-Human and Policy Dimensions 14 (1): 87–99.

Verma, L. R. 1998. Soil and water management techniques. In Indigenous technology knowledge for watershed management in upper north-west Himalayas of India. PWMTA Field Document No. 15. Kathmandu: Participatory Watershed Management Training in Asia (PWMTA) Program. http://www.fao.org/docrep/X5672E/x5672e03.htm.

von Braun, J. 2008. The World Food Situation: New Driving Forces and Required Actions. Food Policy Report, International Food Policy Research Institute, Washington, D.C.

Vorosmarty, C. J., P. Green, J. Salisbury, and R. B. Lammers. 2000. Global water resources: Vulnerability from climate change and population growth. Science 289 (5477): 284–88.

Wang, M., C. Saricks, and D. Santini. 1999. Effects of fuel ethanol use on fuel-cycle energy and greenhouse gas emissions. Argonne, IL: Center for Transportation Research, Energy Systems Division, Argonne National Laboratory.

Warren, R., N. Arnell, R. Nicholls, P. Levy, and J. Price. 2006. Understanding the regional impacts of climate change. Research report prepared for the Stern Review on the Economics of Climate Change. Working Paper No. 90. Norwich, UK: Tyndall Centre for Climate Change Research.

Wassman, R., K. Butterbach-Bahl, and A. Dobermann. 2006. Irrigated rice production systems and greenhouse gas emissions: Crop and residue management trends, climate change impacts, and mitigation strategies. CAB Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources 1 (74): 1–13.

Page 149: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

149

WEDO (Women's Environment and Development Organization). 2007. Changing the climate: Why women's perspectives matter. New York.

WHO (World Health Organization). 2007. Climate and health. Fact Sheet No. 266. Geneva.

WMO (World Meteorological Organization). 2005. Climate and land degradation. WMO Pub No. 989. Geneva.

Woodward, A., S. Hales, and P. Weinstein. 1998. Climate change and human health in the Asia Pacific region: Who will be most vulnerable? Climate Research 11 (1): 31–38.

World Bank. 2000. Cities, Sea, and Storms. Managing Change in Pacific Islands Economies. Volume IV. Adapting to Climate Change. Washington, DC: Papua New Guinea and Pacific Island Country Unit, the World Bank

World Bank. 2000. Cities, sea, and storms. Managing change in Pacific Islands economies. Volume IV, Adapting to climate change. Washington, DC: Papua New Guinea and Pacific Island Country Unit, World Bank.

World Bank. 2000. Cities, sea, and storms: Managing change in Pacific Islands economies. Volume IV, Adapting to climate change. Washington, DC: Papua New Guinea and Pacific Island Country Unit, World Bank.

World Bank. 2003. Piloting weather insurance in India. Washington, DC. August 27. http://go.worldbank.org/FXYEW4DB60.

World Bank. 2007. Global Economic Prospects: Managing the next wave of globalization. The International Bank for Reconstruction and Development, Washington, DC.

World Bank. 2008. World development report 2008: Agriculture for development. International Bank fo Reconstruction and Development, Washington.

WRI (World Resources Institute). 2008. Climate analysis indicators toolkit (CAIT). Available at http://cait.wri.org/.

Wu, N. 1998. Indigenous knowledge of yak breeding and cross-breeding among nomads in western Sichuan, China. Indigenous Knowledge and Development Monitor 6 (1): 7–9

Yari, M. 2003. Beyond "subsistence affluence": Poverty in Pacific Island countries. Bangkok: United Nations Economic and Social Commission for Asia and the Pacific (ESCAP).

Yohe, G., E. Malone, A. Brenkert, M. Schlesinger, H. Meij, and X. Xing. 2006. Global distributions of vulnerability to climate change. Integrated Assessment Journal 6 (3): 35–44.

Yohe, G., E. Malone, A. Brenkert, M. Schlesinger, H. Meij, and X. Xing. 2006. Global Distributions of Vulnerability to Climate Change. The Integrated Assessment Journal 6 (3): 35-44

Yusuf, A. A., and H. Francisco. 2009. Climate change vulnerability mapping for Southeast Asia. South Bridge Court, Singapore: Economy and Environment Program for Southeast Asia (EEPSEA).

Zhou, X. N., G. J. Yang, K. Yang, X. H. Wang, Q. B. Hong, L. P. Sun, J. B. Malone, T. K. Kristensen, N. R. Bergquist, and J. Utzinger. 2008. Potential impact of climate change on

Page 150: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

150

schistosomiasis transmission in China. American Journal of Tropical Medicine and Hygiene 78 (2): 188–94.

Zhu, T. 2007. Review of the impacts of future CO2 and climate on agricultural production. Unpublished.

Page 151: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

151

TABLES

Table 1.1. Climate and agriculture indicators

Country

Climate Variables (1961-1990 average) a

Rural Population

density (people per sq. km of

arable land) - 2002

b

Irrigated land (%

of cropland) 2003-05 c

Agriculture (% of GDP)

c Agricultural Employment

(% of total employment)

2004 d

Dietary Energy Consumption

(kcal/person/day) e

Proportion of undernourished

in total population e

Precipitation (mm)

Temperature (oC 1995 2006 1995-97 2003-05 1995-

97 2003-05

Central Asia Armenia 505.454 6.35385 202 51.2 42 20 10.9 2080 2310 34 21 Azerbaijan 436.82 12.2133 220 69.1 27 7 25.1 2180 2530 27 12 Georgia 959.962 7.35517 280 44.1 52 13 17.8 2250 2480 24 13 Kazakhstan 251.983 5.98001 30 15.7 13 6 16.1 3250 3110 .. .. Kyrgyz Republic 391.205 1.09647 244 73.1 44 33 23.4 2520 3120 13 .. Tajikistan 493.759 3.01356 488 68.2 38 25 31.2 1940 2070 42 34 Turkmenistan 158.402 15.2324 142 89.2 17 20 31.9 2560 2780 9 6 Uzbekistan 192.542 12.5135 357 87.4 32 26 25.0 2710 2440 5 14 East Asia China 571.773 6.28459 559 35.5 20 12 64.4 2840 2990 12 9 Hong Kong, China .. .. .. .. 0 0 .. .. .. .. .. Korea, Rep. 1344.74 10.6524 481 47.1 6 3 7.7 3020 3030 .. .. Mongolia 228.086 -0.497568 88 7 41 22 21.5 1960 2190 40 29 Taipei, China 2005.08 19.0231 .. .. .. .. .. .. .. .. .. Pacific Islands Cook Islands 2182 24 .. .. .. .. 28.6 .. .. .. .. Fiji 2827 24.05 .. .. .. .. 38.1 2770 3010 .. .. Kiribati 1046 27.5 .. .. .. .. 23.9 2810 2830 .. .. Marshall Islands .. .. .. .. .. .. 25.0 .. .. .. .. Micronesia, Federated States of .. .. .. .. .. .. 24.4 .. .. .. .. Nauru .. .. .. .. .. .. 33.3 .. .. .. .. Palau, Republic of .. .. .. .. .. .. 22.2 .. .. .. .. Papua New Guinea 3102.1 25.0696 2007 .. 32 42 72.0 .. .. .. .. Samoa .. .. .. .. .. .. 31.8 2520 2820 .. .. Solomon Islands 3157.09 25.68 .. .. .. .. 71.7 2320 2450 .. .. Timor-Leste, Democratic Republic of 1248.02 24.9667 .. .. .. .. 81.1 2400 2160 .. ..

Page 152: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

152

Country

Climate Variables (1961-1990 average) a

Rural Population

density (people per sq. km of

arable land) - 2002

b

Irrigated land (%

of cropland) 2003-05 c

Agriculture (% of GDP)

c Agricultural Employment

(% of total employment)

2004 d

Dietary Energy Consumption

(kcal/person/day) e

Proportion of undernourished

in total population e

Precipitation (mm)

Temperature (oC 1995 2006 1995-97 2003-05 1995-

97 2003-05

Tonga .. .. .. .. .. .. 31.6 .. .. .. .. Tuvalu .. .. .. .. .. .. 25.0 .. .. .. .. Vanuatu 2699.63 23.7 .. .. .. .. 34.4 2560 2730 .. .. South Asia Afghanistan 311.994 12.9243 273 33.8 .. 36 65.7 .. .. .. .. Bangladesh 2285.82 25.4787 1249 54.3 26 20 51.8 1960 2230 40 27 Bhutan 1833.2 9.40769 .. .. .. .. 93.6 .. .. .. .. India 1082.2 23.9513 466 32.7 26 18 57.8 2380 2360 21 21 Maldives .. .. .. .. .. .. 19.3 2430 2630 .. .. Nepal 1432.29 12.7423 659 47 42 34 93.0 2180 2430 24 15 Pakistan 305.232 19.9923 447 84.2 26 19 45.0 2390 2340 18 23 Sri Lanka 1709.86 26.7409 1588 34.4 23 16 44.3 2260 2360 24 21 Southeast Asia Cambodia 1926.45 26.9452 292 7 48 30 68.6 1860 2160 41 26 Indonesia 2795.42 25.7287 588 12.7 17 13 45.7 2500 2440 13 17 Lao People's Democratic Republic 1764.04 23.2133 480 17.2 56 42 75.8 2090 2300 26 19 Malaysia 2990.04 25.1495 557 4.8 13 9 15.9 2950 2860 .. .. Myanmar 2033.4 22.9983 353 17.9 60 .. 68.9 2050 2380 34 19 Philippines 2322.25 25.3411 559 14.5 22 14 37.1 2380 2470 18 16 Singapore .. .. .. .. .. .. 0.1 .. .. .. .. Thailand 1543.01 26.2632 310 26.6 10 11 53.3 2370 2490 21 17 Viet Nam 1844.77 24.0737 901 33.9 27 20 65.7 2360 2650 21 14

Sources: a) Mitchell et al. (2004, 2005); b) WDI (2005), c) WDI (2008), d) FAOSTAT/PRODSTAT, e) FAO—Food Security Statistics.

Page 153: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

153

Table 1.2. Production of crops and livestock products in ADB DMCs, 2007

Countries Crops - Main products in 2007 Livestock - Main Products in 2007

Central Asia (‘000 mt) (‘000 mt)

Armenia Potatoes 540 Milk 613Tomatoes 250 Cattle meat 43

Azerbaijan Wheat 1,334 Milk 1,301Potatoes 1,178 Cattle meat 76

Georgia Potatoes 175 Milk 734Grapes 93 Cattle meat 49

Kazakhstan Wheat 16,500 Milk 5,007Barley 2,600 Cattle meat 384

Kyrgyz Republic

Potatoes 1,374 Milk 1,192Wheat 709 Cattle meat 92

Tajikistan Potatoes 660 Milk 529Wheat 612 Goat milk, whole, fresh 55

Turkmenistan Wheat 2,700 Milk 1,333Seed cotton 946 Cattle meat 102

Uzbekistan Wheat 5,900 Milk 5,121Seed cotton 3,300 Cattle meat 586

East Asia

China Rice, paddy 187,040 Pig meat 61,150Maize 151,970 Milk 32,820

Hong Kong, China .. .. Korea,

Republic of Rice, paddy 5,960 Milk 2,140Vegetables fresh 3,550 Pig meat 915

Mongolia Potatoes 114 Milk 335Wheat 110 Sheep meat 72

Taipei, China .. .. .. ..Pacific Islands

Cook Islands Roots and Tubers 3 Pig meat 1Coconuts 2 Hen eggs, in shell 0

Fiji Islands Sugar cane 3,200 Milk 58Coconuts 140 Chicken meat 12

Kiribati Coconuts 110 Pig meat 1Roots and Tubers 8 Chicken meat 0

Marshall Islands,

Republic of Coconuts 20 .. ..

Micronesia, Federated States of

Coconuts 41 Pig meat 1

Cassava 12 Cattle meat 0

Nauru Coconuts 2 Pig meat 0Vegetables fresh 1 Hen eggs, in shell 0

Page 154: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

154

Palau, Republic of .. .. ..Papua New

Guinea Oil palm fruit 1,400 Game meat 330Bananas 870 Pig meat 68

Samoa Coconuts 146 Pig meat 4Bananas 23 Milk 2

Solomon Islands

Coconuts 276 Pig meat 2Oil palm fruit 155 Milk 1

Timor-Leste, Democratic Republic of

Maize 63 Pig meat 10

Cassava 50 Chicken meat 2

Tonga Coconuts 59 Pig meat 2Pumpkins, squash and gourds 21 Milk 0

Tuvalu Coconuts 2 Pig meat 0Vegetables fresh 1 Chicken meat 0

Vanuatu Coconuts 322 Milk 3Roots and Tubers 43 Pig meat 3

South Asia

Afghanistan Wheat 3,800 Milk 2,035Vegetables fresh 540 Cattle meat 175

Bangladesh Rice, paddy 43,504 Goat milk, whole, fresh 2,016Sugar cane 6,000 Milk 818

Bhutan Maize 95 Milk 41Rice, paddy 69 Cattle meat 5

India Sugar cane 355,520 Buffalo milk, whole, fresh 56,960Rice, paddy 141,134 Milk 42,140

Maldives Vegetables fresh 28 .. ..Bananas 11 .. ..

Nepal Rice, paddy 3,681 Buffalo milk, whole, fresh 930Sugar cane 2,600 Milk 386

Pakistan Sugar cane 54,752 Buffalo milk, whole, fresh 21,500Wheat 23,520 Milk 11,000

Sri Lanka Rice, paddy 3,131 Milk 143Coconuts 954 Chicken meat 65

Southeast Asia

Cambodia Rice, paddy 5,995 Pig meat 140Cassava 2,000 Cattle meat 63

Indonesia Oil palm fruit 78,000 Poultry Meat + 1,356Rice, paddy 57,049 Chicken meat 1,331

Lao People's Democratic

Republic

Rice, paddy 2,870 Pig meat 47

Vegetables fresh 660 Buffalo meat 19

Malaysia Oil palm fruit 77,700 Chicken meat 931Rice, paddy 2,231 Hen eggs, in shell 465

Myanmar Rice, paddy 32,610 Milk 900

Page 155: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

155

Sugar cane 7,450 Chicken meat 653

Philippines Sugar cane 25,300 Pig meat 1,501Rice, paddy 16,000 Chicken meat 638

Singapore Vegetables fresh nes 18 Chicken meat 76Spinach 2 Hen eggs, in shell 21

Thailand Sugar cane 64,366 Chicken meat 1,050Rice, paddy 27,879 Pig meat 700

Viet Nam Rice, paddy 35,567 Pig meat 2,500Sugar cane 16,000 Chicken meat 344

Source: FAO/PRODSTAT, Accessed April 2009.

Page 156: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

156

Table 1.3. Change in long-term temperature and precipitation under climate change (HadCM3 A2a scenario)

Country

Precipitation (mmH2O/yr) Precipitation

(% change)

Temperature (oC ) Temp (delta

change) Mean_2050 Mean_ Current

t_2050 t_current

Central Asia Armenia 431.2 505.8 -14.7 9.6 5.7 3.9Azerbaijan 398.0 441.6 -9.9 15.7 12.3 3.4Georgia 862.3 980.2 -12.0 10.8 7.2 3.6Kazakhstan 251.6 247.6 1.6 9.1 5.4 3.8Kyrgyz Republic 430.7 421.3 2.2 4.1 0.6 3.6Tajikistan 589.9 559.6 5.4 5.2 1.9 3.3Turkmenistan 169.1 165.9 2.0 17.8 14.8 2.9Uzbekistan 204.2 195.1 4.7 15.1 11.9 3.2

East Asia China 610.9 561.8 8.7 9.0 5.9 3.1Hong Kong, China .. .. .. .. .. .. Korea, Republic of 1642.9 1321.0 24.4 13.8 10.7 3.0Mongolia 204.8 196.6 4.2 2.7 -0.6 3.3Taipei, China 2868.7 2509.2 14.3 20.3 18.6 1.7

Pacific Islands Cook Islands 2468.5 2162.5 14.2 24.0 23.0 1.0Fiji Islands 2934.7 2591.5 13.2 24.9 23.7 1.2Kiribati .. .. .. .. .. .. Marshall Islands, Republic of .. .. .. .. .. .. Micronesia, Federated States of 6068.0 4865.0 24.7 27.0 26.0 1.0Nauru .. .. .. .. .. .. Palau, Republic of 4185.5 3527.5 18.7 28.0 27.0 1.0Papua New Guinea 3469.8 3061.8 13.3 24.9 23.2 1.7Solomon Islands 3623.6 3281.1 10.4 26.5 25.0 1.5Timor-Leste, Democratic Republic of 1595.1 1702.5 -6.3 25.5 23.8 1.6Tonga .. .. .. .. .. ..

Page 157: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

157

Country

Precipitation (mmH2O/yr) Precipitation

(% change)

Temperature (oC ) Temp (delta

change) Mean_2050 Mean_ Current

t_2050 t_current

Tuvalu .. .. .. .. .. .. Vanuatu 2678.6 2627.7 1.9 24.5 23.0 1.5

South Asia Afghanistan, Islamic Republic of 327.3 312.8 4.6 14.3 11.1 3.1Bangladesh 2437.5 2229.2 9.3 27.6 25.0 2.6Bhutan 1632.2 1570.0 4.0 12.1 9.8 2.3India 1210.0 1113.6 8.7 25.9 23.5 2.5Maldives .. .. .. .. .. .. Nepal 1618.6 1371.2 18.0 16.1 13.6 2.5Pakistan 316.7 264.9 19.6 22.6 19.6 3.0Sri Lanka 1894.0 1881.3 0.7 27.9 26.0 1.8

Southeast Asia Cambodia 1674.5 1779.7 -5.9 28.9 26.3 2.6Indonesia 2770.8 2704.9 2.4 26.3 24.6 1.7Lao People's Democratic Republic 1972.7 1888.4 4.5 24.7 22.4 2.3Malaysia 2837.4 2943.8 -3.6 26.9 25.0 1.9Myanmar 2210.1 1989.1 11.1 24.9 22.7 2.3Philippines 2583.6 2543.9 1.6 26.5 24.9 1.6Singapore 2505.0 2473.0 1.3 29.0 26.0 3.0Thailand 1499.4 1497.9 0.1 27.8 25.6 2.2Viet Nam 1808.3 1795.3 0.7 25.4 22.9 2.5

Note: HadCM3 A2a scenario is used for the quantitative analysis in this volume.

Sources: Mitchell et al. (2004, 2005); Hijmans et al. (2005).

Page 158: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

158

Table 1.4: Countries vulnerable to sea level rise and climate extreme events Countries Sea-level rise Floods Droughts Storms

Central Asia

Armenia x x

Azerbaijan x x

Georgia x x x

Kazakhstan x x

Kyrgyz Republic x x

Tajikistan x x

Turkmenistan x

Uzbekistan x x

East Asia

China x x x x

Hong Kong, China x x x

Korea, Rep. x x x x

Mongolia x x x

Taipei, China x x x

Pacific Islands

Cook Islands x x

Fiji x x x x

Kiribati x x x x

Marshall Islands x x x

Micronesia, Federated States of x x x

Nauru (n.a) x

Palau, Republic of (n.a) x

Papua New Guinea x x x x

Samoa x x x

Solomon Islands x x

Timor-Leste, Democratic Republic of

x x x

Tonga x x

Tuvalu x x

Vanuatu x x x

South Asia

Afghanistan x x x

Bangladesh x x x x

Bhutan x x

India x x X x

Maldives x x x

Nepal x X

Pakistan x x X x

Page 159: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

159

Sri Lanka x x X x

Southeast Asia

Cambodia x X

Indonesia x x X x

Lao People's Democratic Republic x X x

Malaysia x x x

Myanmar x x x

Philippines x x X x

Singapore x -

Thailand x x X x

Viet Nam x x X x

Note: Disasters were taken from EM-DAT lists - top ten natural disasters by numbers of people killed, affected people and economic damage costs for the period 1900-2008.

(n.a) - no information available for the country

Source: EM-DAT – Emergency Events Database (for floods, droughts and storms). http://www.emdat.be/Database/CountryProfile/countryprofiles.php (accessed April 20, 2009)

Page 160: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

160

Table 2.1. Summary of studies that assess impacts of climate change on crop yield in Asia

Region/ country

Temperature increments

Scenario/model Crop Yield change with CO2 fertilization

Yield change without CO2

fertilization East Asia China a

1.00 to 1.16 oC (by 2020 above 1990 levels) Precipitation increases by 3.3 to 3.7%

PRECIS (first value: A2 scenarios and second value: B2 scenario)

Maize (irrigated) -0.6 to -0.1% -5.3 to 0.2% Maize (rainfed) +9.8 to +1.1% -10.3 to -11.3% Rice (Irrigated)

+3.8 to -0.4% -8.9 to -1.1%

Rice (rainfed) +2.1 to +0.2% -12.9 to -5.3% Wheat (irrigated) 13.3 to 11.0% -5.6 to -0.5% Wheat (rainfed) +15.4 to +4.5% -18.5 to -10.2%

2.11 to 2.20 oC (by 2050 above 1990 levels) Precipitation increases by 7.0%

PRECIS (first value: A2 scenarios and second value: B2 scenario)

Maize (irrigated) -2.2 to -1.3% -11.9 to -0.4% Maize (rainfed) +18.4 to +8.5% -22.8 to -14.5% Rice (irrigated)

+6.2 to -1.2% -12.4 to -4.3%

Rice (rainfed) +3.4 to -0.9% -13.6 to -8.5% Wheat (irrigated) +25.1 to +14.2% -6.7 to -2.2% Wheat (rainfed) +20.0 to +6.6% -20.4 to -11.4%

3.89 to 3.20 oC (by 2080 above 1990 levels) Precipitation increases by 10.2 to 12.9%

PRECIS (first value: A2 scenarios and second value: B2 scenario)

Maize (irrigated) -2.8 to -2.2% -14.4 to -3.8% Maize (rainfed) +20.3 to 10.4% -36.4 to -26.9% Rice (irrigated)

+7.8 to -4.9% -16.8 to -12.4%

Rice (rainfed) +4.3 to -2.5% -28.6 to -15.7% Wheat (irrigated) +40.3 to +25.5% -8.9 to -8.4% Wheat (rainfed) +23.6 to +12.7 -21.7 to -12.9

Mongoliab

2020 HadCM3 (A2 and B2 scenarios for four ecosystems)

Ewe weight -26.97 to 2.18% 2050 -38.33 to 3.87% 2080 -57.75 to -

0.18% Republic of Korea c

1oC ORYZA 2000 Rice +1.1 to +4.9% -7.50% 2oC -4.4 to -1.1% -12.10% 3oC -7.7 to -4.3% -14.50%

South Asia Islmabad, Pakistan (humid zone) d

0.9oC by 2020 Wheat +8.2%

Faisalabad, Pakistan (semi-arid zone)

0.9oC by 2020 Wheat +2.5%

Multan, Pakistan (arid zone)

0.9oC by 2020 Wheat No effect No effect

Faisalabad, Pakistan (semi-arid zone)

1.8oC by 2050 Wheat -4.0%

Kerala, India e

1.5oC (2040–49 with respect to 1980) 2mm/day for the scenario with CO2 effects

CERES-Rice v3. Rice +5 to +17% (Assuming increase in rainfall by 2mm)

-5 to -14%

Page 161: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

161

and zero increases for scenario without CO2

India (North region) f

1oC Ceres-Rice and ORYZA1N

Rice

+12% (550ppm) -5 % (330ppm) 2.0oC +7% (550ppm) -12% (330ppm) 3.0oC +1% (550ppm) -21% (330ppm) 4.0oC -5% (550ppm) -25% (330ppm) 5.0oC -11% (550ppm) -31% (330ppm)

Bangladesh g

2.0oC CERES-Wheat Model

Wheat

-1% (580ppm), +9% (660ppm)

-37% (330ppm)

4.0oC -40% (580ppm), -31% (660ppm)

-68% (330ppm)

2.0oC CERES - Rice Rice (average of HYV Aus, HYV Aman and HYV Boro)

+14.6% (580ppm), +22.6% (660ppm)

-12% (330ppm)

4.0oC +3.6% (580ppm), +12% (660ppm)

-23% (330ppm)

South East/Mekong Savannakhet Province, Lao PDR h

DSSAT Rice -9.14% (550ppm), -2.56% (720ppm)

Ubonratchathani Province, Thailand (average of 5 zones)

Rice +3.07% (550ppm), +5.96% (720ppm)

An Giang, Can Tho, Dong Thap and Long An Provinces, Viet Nam (average of provinces) Winter

-Spring crop

DSSAT Rice +3.88% (550ppm), -6.95% (720ppm)

Summer-Autumn crop

Rice -5.61% (550ppm), -31.34% (720ppm)

a Erda et al. 2005. b Batima 2006. c Kang Su and Chung Kuen 2006. d O'Brien 2000. e Saseendran et al. 2000. f Aggarwal and Mall 2002. g Karim, Hussain, and Ahmed 1999. h Snidvongs 2006.

Page 162: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

162

Table 2.2. Estimated change in average yield in Viti Levu, Fiji, by 2050

Crop

Change in average yield (%)

Impact of change in average rainfall and

temperature

Impact of change in rainfall, temperature, and climate variability (ENSO)

Sugarcane n.a -9Dalo (Taro) -5 to +1 -15 to +1Yam -5 to +4 -11 to +4Cassava -9 to -5 -12 to -6

Source: World Bank 2000.

Page 163: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

163

Table 2.3. Difference in 2050 Baseline and Climate Change Scenario Production.

Central Asia East Asia China

South Asia

India Southeast

Asia

Maize -5% -10% -10% -19% -56% 27%

Rice 2% 10% 10% 5% 1% -2%

Wheat -11% 18% 18% 17% 14% -1%

Page 164: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

164

Table 2.4. Difference in 2050 Baseline and Climate Change Scenario Area.

Central Asia East Asia China

South Asia

India Southeast

Asia

Maize -20% -1% -1% -18% -35% 12%

Rice -3% 4% 4% -14% -25% -2%

Wheat 0% 13% 13% 1% 0% 10%

Page 165: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

165

Table 2.5. Difference in 2050 Baseline and Climate Change Scenario Cereal Yields.

Central Asia East Asia China

South Asia

India Southeast

Asia

Maize 19% -10% -10% -1% -32% 14%

Rice 5% 6% 6% 22% 35% 0%

Wheat -11% 4% 4% 16% 14% -10%

Page 166: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

166

Table 2.6. Impacts of climate change on agriculture GDP of Asian countries by 2100

Country

AGR GDPa

(billion US$/year)

Cross-sectional response function (billion

US$/year)

Experimental climate response function (billion

US$/year) CCC CCSR PCM CCC CCSR PCM

China 230.5 -5.6 -2.0 7.6 -32.2 7.0 30.5Bangladesh 16 -0.1 -0.1 -0.1 -9.4 -3.7 -0.5India 164.9 -1.1 -1.5 0.5 -86.7 -47.8 -4.6Nepal 3.4 -0.0 -0.0 0.0 -0.5 -0.5 0.3Pakistan 19.3 -0.2 -0.2 0.5 -15.8 0.5 1.8Sri Lanka 3.4 -0.0 0.0 0.0 -1.8 -0.6 0.4Cambodia 0.2 -0.0 -0.0 -0.0 -0.2 -0.2 1.1Indonesia 38.4 -0.3 -0.2 -0.1 -25.5 3.3 -4.3Laos 0.2 -0.0 -0.0 0.0 -0.2 -0.2 0.2Malaysia 4.2 -0.4 0.1 0.3 -5.9 -0.4 4.2Myanmar 11.5 -0.2 -0.2 0.0 -1.1 -2.7 1.3Philippines 14.6 -0.5 -0.1 0.4 -8.8 -2.7 1.0Thailand 19.6 -1.5 -1.6 -0.3 -19.6 -11.1 3.7Viet Nam 4 -0.1 -0.1 0.1 -4.0 -2.2 0.7a Total projected value of agricultural GDP in 2100 at current climate (temperature: 23.7oC; precipitation: 15.2 millimeters a month [mm/mo]). Note: CCC: temperature:28.1oC; precipitation: 13.6 mm/mo. CCSR: temperature: 27.0oC; precipitation: 16.4 mm/mo. PCM: temperature: 25.9oC; precipitation: 18.1 mm/mo.

Source: Mendelsohn 2005.

Page 167: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

167

Table 2.7. Per capita consumption (kg per year) of cereal and meat with and without climate change.

No Climate

Change SRES A2a

2000 2050 2050 Cereal East Asia 182 154 144

China 184 155 145South Asia 163 157 144India 161 155 142Southeast Asia 182 148 139Developed 116 120 107

Meat East Asia 49 96 91China 50 98 93South Asia 6 21 20India 5 21 20Southeast Asia 18 44 42Developed 86 91 87

 

Page 168: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

168

Table 2.8. Prevalence of undernourishment in Asia and Pacific countries

Region/Country

Total population

2003-05 (millions)

Number of people undernourished (millions)

Proportion of undernourished in total population

1990-92 1995-97 2003-

05 1990-

92 1995-97 2003-05

East Asia 1362.7 178.7 144.7 123.5

China 1312.4 178 143.7 122.7 15 12 9

Mongolia 2.6 0.7 1 0.8 30 40 29

Republic of Korea 47.7 - - - - - -

Southeast Asia 544.6 105.2 88.3 86.3

Cambodia 13.7 3.8 4.8 3.6 38 41 26

Indonesia 223.2 34.5 26.7 37.1 19 13 17 Lao People's Dem Rep. 5.6 1.1 1.3 1.1 27 26 19

Malaysia 25.2 - - - - - -

Myanmar 47.6 18.1 14.8 8.8 44 34 19

Philippines 82.9 13.3 12.8 13.3 21 18 16

Thailand 62.6 15.7 12.3 10.9 29 21 17

Viet Nam 83.8 18.7 15.6 11.5 28 21 14

South Asia 1468.5 282.5 284.7 313.6

Bangladesh 150.5 41.6 51.4 40.1 36 40 27

India 1117 206.6 199.9 230.5 24 21 21

Nepal 26.6 4 5.3 4 21 24 15

Pakistan 155.4 25.7 23.7 35 22 18 23

Sri Lanka 19 4.6 4.4 4 27 24 21

Central Asia 73.6 10 8.9 8.3

Kazakhstan 15.1 - - - - - -

Kyrgyzstan 5.2 0.8 0.6 - 17 13 -

Tajikistan 6.5 1.8 2.4 2.2 34 42 34

Turkmenistan 4.8 0.3 0.4 0.3 9 9 6

Uzbekistan 26.2 1 1.1 3.6 5 5 14

Armenia 3 1.6 1.1 0.6 46 34 21

Azerbaijan 8.3 2 2.1 1 27 27 12

Georgia 4.5 2.5 1.2 0.6 47 24 13

Source: FAO (2008e)

Page 169: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

169

Table 2.9. Number of malnourished children in developing Asia under Climate Change (thousands of children, under 5 years of age).

2000 2050

no CC Additional with CC Afghanistan 876 978 224

Bangladesh 9055 7963 158

Bhutan 22 20 3

China 9586 1173 676

India 56431 34540 983

Indonesia 5323 3253 163

Kazakhstan 48 37 11

Kyrgistan 31 8 9

Malaysia 520 261 16

Mongolia 34 9 4

Myanmar 957 661 10

Nepal 1760 1372 80

Pakistan 7162 5176 197

Papua New Guinea 159 78 5

Philippines 1889 908 5

South Korea 591 318 2

SouthEast Asia 1085 1266 5

Sri Lanka 316 175 10

Tajikistan 167 52 14

Thailand 968 641 16

Turkmenistan 58 38 8

Uzbekhistan 540 395 35

Viet Nam 2754 1901 65

Source: IMPACT model simulations, A2a SRES scenario

Page 170: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

170

Table 2.10. Top 25 Natural Disasters by number of total affected people in 2006

Country Disaster Disaster name Number of affected people1 China Typhoon Bilis 29,622,000 2 China Drought 18,000,000 3 China Typhoon Prapiroon 10,000,000 4 China Cyclone Kaemi 6,531,000 5 India Flood 6,000,065 6 China Typhoon 5,920,000 7 China Flood 4,600,024 8 Malawi Drought 4,500,000 9 China Flood 4,120,000

10 Philippines Typhoon Xangsane (Milenyo) 3,842,406 11 Kenya Drought 3,500,000 12 Indonesia Earthquake 3,177,923 13 China Tropical storm Chanchu (Caloy) 3,150,000

14 Philippines Typhoon Durian (Reming) 2,562,517 15 China Valley flood 2,375,000 16 Thailand Flash flood 2,212,413 17 Burundi Drought 2,150,000 18 India Flood 2,000,000 19 Afghanistan Drought 1,900,000 20 Viet Nam Typhoon Xangsane (Milenyo) 1,467,925 21 China Flood 1,410,000 22 China Flood 1,400,000

23 Viet Nam Typhoon Durian (Reming) 1,226,360 24 Kenya Flood 723,000 25 Thailand Flash flood 700,000

Source: Sanker, Nakano, and Shiomi 2007

Page 171: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

171

Table 3.1. Disaster and impacts of hydro-meteorological disasters by income level, 1975-2006 (world %)

Income class Total of people killed Total of people

affected Total of damage

High Income 4.4 0.8 55.4

Lower Income 79.2 52.0 7.9 Lower Middle Income 11.7 45.5 30.4 Upper Middle Income 4.65 1.7 6.3

Source: Sanker, Nakano, and Shiomi (2007).

Page 172: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

172

Table 3.2 Results from previous vulnerability assessments

Scenario Vulnerability 2050 2100 Source and methodology

Static index of current national adaptive capacities

Extreme Bangladesh All countries investigated Vulnerability as a function of exposure,

sensitivity, and adaptive capacity. Assumption that the climate sensitivity is 5.5°C. Scenario: A2. Source: Yohe et al.

(2006). Observations: i) Countries not in the sample: Central Asia,

Myanmar, Afghanistan, Laos, most of Pacific Island countries, Mongolia. ii)PNG - Papua

New Guinea

Significant China, Pakistan, Nepal -

Moderate India, Southeast Asia, PNG, Sri Lanka -

Little or no - -

National adaptive capacities improving over time

Extreme -

China, India, Pakistan, PNG, Nepal, Bangladesh, Viet Nam, Cambodia, Philippines

Significant - Other Southeast Asia, Sri Lanka

Moderate China, India, Pakistan, PNG, Nepal, Bangladesh -

Little or no Southeast Asia, Sri Lanka -

Mitigation scenario (greenhouse gas concentration constrained to 550 ppm) combined with a scenario of static adaptive capacity

Extreme - China, India, Pakistan, Nepal, Bangladesh, Cambodia

Significant Nepal, Bangladesh Other Southeast Asia, PNG, Sri Lanka

Moderate China, India, Pakistan, PNG, Cambodia -

Little or no Other Southeast Asia, Sri Lanka -

Mitigation scenario combined with an enhanced adaptive capacities scenario

Extreme - China, Pakistan, Nepal

Significant - India, PNG, Bangladesh, Viet Nam, Philippines

Moderate China, India, Pakistan, Nepal, Bangladesh Other Southeast Asia, Sri Lanka

Little or no Southeast Asia, PNG, Sri Lanka -

Page 173: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

173

Scenario Vulnerability - Resilience

Indicators 2095 1990 Source and methodology

Current Resilient countries Indonesia, Republic of Korea, Cambodia

38 countries around the world were investigated and eight in the

ADB region. Vulnerability to climate change is assessed by

combining indicators that measure sensitivity to climate

change (settlement, food security, human health, ecosystems, and

water indicator) and coping-adaptive capacity (economic

capacity, human resources, and environmental capacity) A

vulnerability-resilience indicator is calculated by estimating the difference between aggregated

sensitivity and adaptation capacity. Exposure was not

assessed in this study. Source: Moss et al. (2001)

Vulnerable countries Uzbekistan, Thailand, Bangladesh, India, China

A1v2 (rapid growth) Resilient countries

Indonesia, Republic of Korea, Thailand, Cambodia, China, Uzbekistan, Bangladesh, India

Vulnerable countries -

B2h (local sustainability)

Resilient countries Indonesia, Republic of Korea, Thailand, Cambodia, China, Uzbekistan

Vulnerable countries India, Bangladesh

A2A1 (delayed development)

Resilient countries Indonesia, Republic of Korea, Thailand, Cambodia

Vulnerable countries India, Bangladesh, Uzbekistan, China

Time/Scenario Vulnerability (to mortality from climate related disasters) Source and methodology

Current

Most vulnerable countries Afghanistan, Pakistan, Turkmenistan,

This study assesses vulnerability to climate-related mortality at the national level. Climate outcomes are represented by mortality from climate-related disasters using the emergency events database data set, statistical relationships between mortality and a shortlist of potential proxies for vulnerability are used to identify key vulnerability indicators.. Eleven key indicators related to health status, governance and education were found exhibit a strong relationship with decadally aggregated mortality associated with climate-related disasters. Validation of indicators, relationships between vulnerability and adaptive capacity, and the sensitivity of subsequent vulnerability assessments to different sets of weightings are explored using expert judgement data, collected through a focus group exercise. Source: Brooks et al. (2005) Moderately to highly vulnerable

Laos, Tonga, Nepal, Azerbaijan, Bangladesh, Bhutan, Cambodia,

Page 174: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

174

Time/scenario Human Vulnerability to Climate Extremes Source and methodology

Next 20-30 years (GIS)

High Human Vulnerability

Afghanistan, Pakistan, India, Myanmar, Laos, Cambodia and Indonesia, Mongolia and northern and western China

The authors use Geographical Information Systems (GIS) to map specific hazards associated with climate change – specifically: floods,

cyclones and droughts – and place them in relation to factors influencing vulnerability. This study identifies the most likely

humanitarian implications of climate change for the next 20-30 year period. In this study, human vulnerability refers to the likelihood that individuals, communities or societies will be harmed by a hazard. We

divide the factors shaping human vulnerability into five groups: natural, human, social, financial and physical. Each group contains one or more individual indicators, which were combined to construct vulnerability

indices for each. The groups were then combined to give a single, overall human vulnerability index. This was combined with

information on the distribution of hazards to identify climate change-risk hotspots. Source: Ehrhart et al. (2008)

Flood risk hotspots South and Southeast Asia

Cyclone risk hotspots

Bangladesh, several parts of India, Viet Nam and several other Southeast Asian countries

Drought risk hotspots

Afghanistan, Pakistan and parts of India, Myanmar, Viet Nam and Indonesia

Climate extreme risk hotspots + high population density South and Southeast Asia

Drought risk hotspots + risk of conflict South Asia

Time/scenario Vulnerability to climate hazards Source and Methodology Current (only in Southeast Asia)

Overall vulnerability

All the regions of the Philippines, the Mekong River Delta region of Viet Nam; almost all the regions of Cambodia; North and East Lao PDR; the Bangkok region of Thailand; and the west and south of Sumatra, and western and eastern Java in Indonesia

Vulnerability to climate change in Southeast Asia is measured by taking into account indicators of exposure (multiple hazard risk

exposure), sensitivity (human and ecological), and adaptive capacity. Authors assessed exposure using information from historical records of climate-related hazards (tropical cyclones, floods, landslides, droughts, and sea level rise). Population density was used as a proxy for human

sensitivity to climate-hazard exposure. Biodiversity information (percentage of protected areas) was used as a proxy to ecological

sensitivity. An index of adaptive capacity was created as a function of socio-economic factors, technology, and infrastructure. Source: Yusuf

and Francisco (2009).

Drought hotspots

Northwestern Viet Nam, Eastern coastal areas of Viet Nam, Southern regions of Thailand, the Philippines, Sabah state in Malaysia, Western and eastern area of Java Island, Indonesia

Page 175: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

175

Floods hotspots

Bangkok and its surrounding area in Thailand, Southern regions of Thailand, the Philippines, Western and eastern area of Java Island, Indonesia

Sea level rise hotspots

Mekong region of Viet Nam, Bangkok and its surrounding area in Thailand, Western and eastern area of Java Island, Indonesia

Cyclone hotspots Eastern coastal areas of Viet Nam, the Philippines

Landslide hotspots

the Philippines, Western and eastern area of Java Island, Indonesia

Note: Yohe et al. (2006) estimate vulnerability considering several assumptions: a static index of current national adaptive capacities (Panel A of figure 3.2); with national adaptive capacities improving over time (Panel B); under a mitigation scenario (greenhouse gases concentration constrained to 550 ppm) combined with a scenario of static adaptive capacity (Panel C); and under a mitigation scenario combined with an enhanced adaptive capacities scenario (Panel D). Moss, Brenkert and Malone (2001) assess vulnerability to climate change by combining indicators that measure sensitivity to climate change (settlement, food security, human health, ecosystems, and water indicator) and coping-adaptive capacity (economic capacity, human resources, and environmental capacity) for 1990 and 2095. A vulnerability-resilience indicator is calculated by estimating the difference between aggregated sensitivity and adaptation capacity. By 1990, 16 out of 38 countries were considered vulnerable to climate change including Uzbekistan, Thailand, Bangladesh, China and India. By the year 2095, only one country remains vulnerable in the rapid growth scenario -A1v2 (Yemen), three countries in the local sustainability scenario - B2h (Yemen, India and Bangladesh) and nine countries in the delayed development scenario - A2A1 (Yemen, India, Bangladesh, Ukraine, Uzbekistan, Tunisia, China, Egypt, South Africa, Senegal). Brooks, Adger and Kelly (2005) find that 11 key indicators related to health status, governance and education exhibit a strong relationship with decadally aggregated mortality associated with climate-related disasters. Through a focus group exercise, expert judgment data was used for validation of indicators, relationships between vulnerability and adaptive capacity, and the sensitivity of subsequent vulnerability assessments to different sets of weightings. Following this methodology, most vulnerable countries were found to be mainly in Sub-Saharan Africa. However, Afghanistan, Pakistan and Turkmenistan were also in the most vulnerable group..

Page 176: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

176

Table 3.3 Countries identified as vulnerability to climate change in Asia

Note: Poor outcomes in all three areas, shaded in dark grey, indicate high vulnerability, and poor outcomes in two areas, shaded in light grey, indicate significant vulnerability. Adaptive capacity was represented as level of poverty (Bauer et al. 2008). A poverty level of more than 30 percent is considered as low adaptive capacity. Sensitivity was represented by share of labor employed in agriculture (FAO, 2004); countries with agricultural employment above 40 percent are considered highly sensitive. Exposure, finally, was reflected as the delta change in both temperature and annual precipitation in 2080 as compared to current climate—1961-1990. Countries were classified as highly exposed if the temperature increases by at least 2 degrees or if annual precipitation levels increase or decrease by at least 20 percent. The climate scenarios are derived from Hijmans et al. (2005) for the Hadley SRES A2a scenario.

Page 177: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

177

Table 4.1. GHG emissions by main sources in agricultural sector in different regions, 2005 Region N2O

soils CH4

enteric CH4 rice

CH4, N2O manure

CH4, N2O burning

Total

South Asia Mt CO2-eq/yr 536 275 129 40 24 1,005 percent of region’s total 53 27 13 4 4 100 percent of source’s world total

20 15 20 9 3 17

East Asia Mt CO2-eq/yr 600 294 432 127 53 1,505 percent of region’s total 40 20 29 8 4 100 percent of source’s world total

23 16 68 29 14 25

Subtotal (global developing regions)

Mt CO2-eq/yr 1,946 1,300 617 211 363 4,438 percent of region’s total 44 29 14 5 8 100 percent of source’s world total

74 70 97 48 92 74

Subtotal (global developed regions)

Mt CO2-eq/yr 700 554 20 225 32 1,531 percent of region’s total 46 36 1 15 2 100 percent of source’s world total

26 30 3 52 8 26

TOTAL Mt CO2-eq/yr 2,646 1854 637 436 395 5,969 percent of region’s total 44 31 11 7 7 100 percent of source’s world total

100 100 100 100 100 100

Source: USEPA (2006).

Page 178: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

178

Table 4.2. Technical reduction potential (proportion of an animal’s enteric methane production) for enteric methane emissions due to(1i) improved feeding practices, (2) specific agents and dietary additives, and (3) longer-term structural/management change and animal breeding

Regions Improved feeding practices Specific agents and dietary additives Longer-term structural/management

change and animal breeding Dair

y cows

Beef cattle

Sheep

Dairy buffalo

Non-dairy buffalo

Dairy cows

Beef cattle

Sheep

Dairy buffalo

Non-dairy buffalo

Dairy cows

Beef cattle

Sheep

Dairy buffalo

Non-dairy buffalo

South Asia

0.04 0.02 0.02 0.04 0.02 0.01 0.01 0.0005

0.01 0.002 0.01 0.01 0.001 0.01 0.02

East Asia 0.10 0.05 0.03 0.10 0.05 0.03 0.05 0.002 0.03 0.012 0.03 0.06 0.003 0.03 0.07 West Asia

0.06 0.03 0.02 0.06 0.03 0.01 0.02 0.001 0.01 0.004 0.01 0.02 0.001 0.02 0.03

Southeast Asia

0.06 0.03 0.02 0.06 0.03 0.01 0.02 0.001 0.01 0.004 0.01 0.02 0.001 0.02 0.03

Central Asia

0.06 0.03 0.02 0.06 0.03 0.01 0.02 0.001 0.01 0.004 0.01 0.02 0.001 0.02 0.03

Source: Smith et al. (2007a).

Page 179: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

179

Table 4.3. Costs of mitigation in subregions of Asia at various carbon prices. Region Potential (Mt CO2-eq. yr-1) Potential (Mt CO2-eq. yr-1) Potential (Mt CO2-eq. yr-1) up to 20 USD t CO2-eq. up to 50 USD t CO2-eq. up to 100 USD t CO2-eq.

Western Asia 84.19 117.67 139.40 Southeast Asia 93.36 190.24 344.38 South Asia 16.40 41.25 84.11 East Asia 27.69 36.67 39.24 Central Asia 55.15 126.61 245.44 Total 276.79 512.44 852.57 Total investments @ carbon price (million USD) 5,535 25,622 85,257

Source: Smith (2009).

Page 180: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

180

Table 5.1: Adaptation and complementary mitigation measures and implementers. Implementer Scope for

mitigation Institutional

support Existing experience in Asia-

Pacific Farm Community National Global

1. Innovative adaptation to climate change

Changes in agricultural practices

Zero till, no till, changing planting dates, crop varieties, soil and water conservation techniques, integrated pest and pathogen management techniques, supplementary livestock feeds

Integrated pest management

Strong extension services, functioning credit markets, market information system, climate information system

Integration of agriculture into carbon markets, efficient quarantine capabilities and monitoring programs

Sustainable land management, pasture and grazing land management, restoration of degraded soils, livestock management, agroforestry practices, biofuels, nutrient management programs, promotion of no-till agriculture

Functioning markets and information services facilitating efficient use of resources and access to information

Alternative wetting and drying for rice production (China); zero/low till (Indo-Gangetic Plains); N fertilizer added to soils only after soil N testing (China); shorter-cycle rice seeds in the Mekong Delta (Viet Nam)

Changes in agricultural water management

Water harvesting, on-farm irrigation, soil and water conservation, drip and sprinkler systems, groundwater use, treadle pumps

Small reservoirs, watershed management, water trading

Investment in large-scale systems, reservoirs,

Support for funding

Water quality legislation limiting nonpoint source pollution from agriculture; groundwater development might increase CO2 emissions as might

Credit availability, monitoring and enforcement of water quality standards

Small tanks for dry-season irrigation (Tamil Nadu/Sri Lanka); pilot water trading (northern China), carbon credits for treadle pumps (India); water quality programs that control nonpoint source pollution (China)

Page 181: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

181

Implementer Scope for mitigation

Institutional support

Existing experience in Asia-Pacific Farm Community National Global

irrigation expansion

Agricultural diversification

Shrimp/rice farming in coastal areas with sea-level rise, migration, off-farm work

Strong extension services

Flexible water control structures supporting shrimp and rice production (Viet Nam’s Mekong Delta)

Agricultural science and technology development

Participatory crop breeding

Drought/heat-resistant crops; salinity tolerant varieties, water-conserving crops, animal breeds

Support for funding, strengthening research for sustained agricultural biodiversity

Breeding for higher fertilizer use efficiency

Capacity building of scientists (curricula at BSc, MSc, PhD level; on-the-job training), public-private partnerships, appropriate IPR/biosafety systems, strong IARCs and NARESs

Aerobic rice (northern China)

Agricultural advisory service and information systems

Farmer-to-farmer training

Support dissemination of climate-resilient varieties, technologies, and practices; disseminate (seasonal) climate

Link farmers to carbon markets, disseminate information on mitigation technologies and practices, capacity

Pluralistic, demand-driven, decentralized advisory service; strong research-extension-farmer

Page 182: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

182

Implementer Scope for mitigation

Institutional support

Existing experience in Asia-Pacific Farm Community National Global

forecasts

building on certification, monitoring, and obtaining carbon credits

linkages

Risk management and crop insurance

Crop insurance

Contract farming; weather index insurance, futures and option contracts

Weather index insurance

Global climate insurance

Appropriate agricultural advisory service, hydro-meteorological infrastructure, functioning financial markets and institutions

Various type of contract farming arrangements for crab and shrimp farmers (Indonesia)

2. Strengthening important ongoing development initiatives to support climate change adaptation

Agricultural market development

Cooperatives for farm inputs and outputs, ICTs

Infrastructure and market development, one-stop information centers, extension service

Carbon market development, carbon market information,

support for farmers to access value chains

Public-private partnerships for carbon markets and new market opportunities as a response to global warming, property rights to land

Agricultural policies Mainstream climate change in policies, phase out fertilizer

Mainstream climate change in global policies, climate

Mainstream win-win mitigation-adaptation policies

Capacity building at the government agency and NARES level

Regional agricultural development programs (China)

Page 183: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

183

Implementer Scope for mitigation

Institutional support

Existing experience in Asia-Pacific Farm Community National Global

subsidies

policies targeting the poor

through appropriate incentives, policies to ameliorate adverse impacts from massive expansion of livestock production (monitoring, regulatory, research, and extension)

Trade Cooperative storage, market information

Trade reform, legislation on food safety standards, support for market information systems, road development, phase out subsidies for biofuel promotion, support research in second-or third-generation technologies

Doha Round of WTO, food safety standards

Focus on mitigation instruments that support trade

Codex Alimentarius, producer organizations, consumer organizations

Other environmental policies

Environmental legislation, including afforestation,

Air quality legislation (reducing straw

Reforestation in northern China

Page 184: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

184

Implementer Scope for mitigation

Institutional support

Existing experience in Asia-Pacific Farm Community National Global

reforestation; secure property rights for land and water; improved incentives and markets in natural resources; policies to maintain ecosystem services

burning); afforestation, reforestation policies; secure property rights to secure carbon financing for mitigation

Enhanced social protection and microfinance

Effective rural institutions, microcredit groups for women

Food reserves and storage, social policies (safety nets) targeted to the poor

Timely, targeted food aid when needed (mostly disaster); government-sponsored food- or cash-for-work programs

Functioning financial markets and institutions (e.g., agricultural credit providers), rules and regulations, decentralization, subsidiarity principle

BRAC for facilitating microcredit (Bangladesh)

Disaster preparedness

Climate information system, early warning systems

Support to investments in these systems, international regulations

Tsunami information system; Bangladesh

Page 185: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

185

Table 5.2. Examples of adaptation measures by sector

Sector Examples of adaptation measures

Water

Groundwater and rainwater harvesting Increased desalination Protection of water catchment areas Improved systems of water management Development of flood controls and drought monitoring Development of early warning systems

Agriculture and food security

Changes in agricultural practices (planting and harvesting times, fertilizer use, pest control, etc.) Improved irrigation techniques Diversification of crops and income sources Development of tolerant crop varieties Improved extension services

Infrastructure / settlement (including coastal zones)

Strengthening of coastal defenses Improvement of key coastal infrastructure and human settlements Integrated coastal zone management Improved coastal planning and land use legislation Support for the relocation of high-risk populations

Human health Improved disease surveillance systems Development of early warning systems Improved preparedness and emergency response

Ecosystems (terrestrial and marine)

Improvement of natural resource management systems Protection of coral reefs and coastal vegetation Improved species monitoring and identification Creation of protected areas and biodiversity corridors Development and maintenance of seed banks

Source: IPCC 2006a.

Page 186: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

186

Table 5.3. Adaptation responses and issues

Type of response

Autonomous Policy driven

Short run Crop choice, crop area, planting date

Risk-pooling insurance

Improved forecasting

Research for improved understanding of climate risk

Long run Private investment (on-farm irrigation)

Private crop research

Large-scale public investment (water, storage, roads)

Crop research

Issues Costly to poor

Social safety nets

Trade-offs with integration

Uncertain returns to investment

Costs

Source: Rosegrant et al. 2008.

Page 187: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

187

Table 5.4. Costs of priority activities of adaptation in selected LDCs in the Asia-Pacific region .

Country Adaptation measure Cost

(US$ millions)

Bangladesh

Construction of flood shelters, and information and assistance centers to cope with more frequent and intense floods in major floodplains

5.00

Enhancing the resilience of urban infrastructure and industries to the impacts of climate change

2.00

Promoting adaptation of coastal crop agriculture to salinity 6.50

Adaptation of fisheries in areas prone to enhanced flooding in the Northeast and Central Regions through adaptive and diversified fish culture practices

4.50

Bhutan

Landslide management and flood prevention 0.89

Weather forecasting system to serve farmers 0.42

Flood protection of downstream industrial and agricultural areas 0.45

Rainwater harvesting 0.90

Cambodia

Rehabilitation of upper Mekong and provincial waterways to reduce risks caused by floods, improve fishery resources and supply sufficient water for irrigation and domestic uses

30.00

Vegetation planning for flood and windstorm protection 4.00

Development and improvement of community irrigation systems 4.00

Community mangrove restoration and sustainable use of natural resources 1.00

Samoa

Reforestation, rehabilitation and community forestry fire prevention project 0.42

Climate early warning system project to implement effective early warnings and emergency response measures to climate and extreme events

4.50

Coastal infrastructure management plans for highly vulnerable districts 0.45

Sustainable tourism that takes into account climate change and climate variability 0.25

Tuvalu

Increasing resilience of coastal areas and settlement to climate change 1.90

Increasing pit-grown pulaka productivity through introduction of a salt-tolerant pulaka species

2.20

Adaptation to frequent water shortages through increasing household water capacity, water collection accessories, and water conservation techniques

2.70

Source: Adapted from NAPAs submitted to the UNFCCC.

Source: IGES 2008.

Page 188: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

188

Table 5.5. The impact of coastal protection on sea-level rise damage (number of people at risk from a one-meter rise in sea level) GVA-case countries People at riska

without measures (‘000 people)

People at riska with additional measures (‘000 people)

Cost of measures (percent of

GNP per year)b North America 170 90 0.02 Central America 56 6 0.23 Caribbean islands 110 20 0.21 South America, Atlantic coast 410 48 0.25 South America, Pacific coast 100 11 0.01 North and West Europe 130 130 0.02 North Mediterranean 37 31 0.02 South Mediterranean 2,100 250 0.07 Africa, Atlantic coast 2,000 220 0.25 Gulf States 14 3 0.05 Asia, Indian Ocean coast 27,360 3,040 0.52 Indian Ocean, small islands 100 12 0.72 Southeast Asia 7,800 880 0.20 East Asia 17,100 2,200 0.06 Pacific Ocean, large islands 17 4 0.17 Pacific Ocean, small islands 34 4 0.77

World 61,300 7,380 0.056 (average)

a. Number of people living in the risk zone, multiplied by the probability of flooding per year.

b. Undiscounted, assuming 100 years lifetime, that is, annual cost is 1 percent of total cost.

Source: Frankhauser et al. (1998) based on IPCC (1994) and Delft Hydraulics (1993) cited in Frankhauser (2006), Francisco (2008).

Page 189: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

189

FIGURES

Figure 1.1. Annual mean temperature change (oCelsius) in 2050s relative to 1950-2000 historical mean (HadCM3/A2a)

Source: Hijmans et al. (2005).

Page 190: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

190

Figure 1.2. Annual mean total precipitation change (mm) in 2050s relative to the 1950-2000 historical mean (HadCM3/A2a)

Source: Hijmans et al. (2005).

Page 191: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

191

Figure 1.3. Climate change vulnerability, adaptation, and resilience

Source: Adapted from Gbetibuou (2009).

EXPOSURE

CharacteristicsFrequency,magnitude,duration

IMPACTS

SENSITIVITY

CharacteristicsAssets, entitlements,economic structures,human capital

DeterminantsCoping strategies, social networks, resource use, diversity and flexibility

ADAPTIVE CAPACITY

VULNERABILITY

ADAPTATION (&MITI-GATION) RESPONSES

RESILIENCE

Page 192: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

192

Figure 2.1. Impact of climate change on cereal production in climate model projections, 2080 (% change from respective reference projection) considering economic adjustments

Source: Fischer Shah, and Van Velthuizen. 2002.

‐15

‐10

‐5

0

5

10

15

A1F1 A2 B2 B1 A2 B2 B1 A2 B2 A2 B2

HadCM3 CSIRO CGCM2 NCAR

World Developed Regions Developing Regions Asia

Page 193: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

193

Figure 2.2 Impact of global warming on South Asia countries' agricultural production capacity by the 2080s (%)

Source: Cline 2007.

-45-40-35-30-25-20-15-10

-50

Bangladesh Nepal Sri Lanka Pakistan Afghanistan India

Without carbon fertilization With carbon fertilization

Page 194: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

194

Figure 2.3. Impact of global warming on agricultural production capacity of regions of China by 2080 (%)

Source: Cline 2007.

-30-25-20-15-10

-505

10152025

Beijing Northeast

Central Hong Kong Southeast

Northwest South Central

Tibetan Plateau

Yellow Sea

Without carbon fertilization With carbon fertilization

Page 195: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

195

Figure 2.4. Impacts of global warming on Southeast Asian countries’ agricultural production capacity by the 2080s (%)

Source: Cline 2007.  

‐45

‐40

‐35

‐30

‐25

‐20

‐15

‐10

‐5

0

Cambodia Indonesia Malaysia Myanmar Philippines Thailand Viet Nam

Without carbon fertilization With carbon fertilization

Page 196: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

196

Figure 2.5. Impact of climate change on direct human cereal consumption, world and Southeast Asia in various models, 2080 (% changes from respective reference projection)

Source: Fischer et al. 2002.

‐5

‐4

‐3

‐2

‐1

0

1

A1F1 A2 B2 B1 A2 B2 B1 A2 B2 A2 B2

HadCM3 CSIRO CGCM2 NCARWorld

South East Asia

Page 197: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

197

Figure 2.6. Impact of global warming on Central Asian countries’ production capacity by the 2080s (%)

Source: Cline 2007.

-15-10

-505

101520253035

Kazakhstan Other Central Asia Uzbekistan

Without carbon fertilization With carbon fertilization

Page 198: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

198

Figure 2.7: Results from the GCM3.1 (T63) model, A1B scenario

Source: http://ipcc-wg1.ucar.edu/wg1/Report/suppl/Ch10/Ch10_indiv-maps.html

Page 199: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

199

Figure 2.8. Results from the Hadley model, A1B scenario

Source: http://ipcc-wg1.ucar.edu/wg1/Report/suppl/Ch10/Ch10_indiv-maps.html

Page 200: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

200

Figure 2.9: Rice, All-Asia Average Response to Temperature and Nitrogen by Rainfed and Irrigated Crops

Figure 2.10: Maize, All-Asia Average Response to Temperature and Nitrogen by Rainfed and Irrigated Crops

Figure 2.11. Rice, China Average Response to Temperature and Nitrogen by Rainfed and Irrigated Crops

Figure 2.12. Rice, India Average Response to Temperature and Nitrogen by Rainfed and Irrigated Crops

Page 201: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

201

Figure 2.13. Rice All-Asia Average Response to CO2 Fertilization and Nitrogen by Rainfed and Irrigated Crops

Figure 2.14. Maize All-Asia Average Response to CO2 Fertilization and Nitrogen by Rainfed and Irrigated Crops

Page 202: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

202

Low input rainfed maize, East Asia High input rainfed maize, East Asia

Low input irrigated rice, East Asia

High input irrigated rice, East Asia

Low input rainfed wheat, East Asia

High input rainfed wheat, East Asia

High input rainfed rice, East Asia

Page 203: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

203

Low input rainfed maize, South Asia

High input rainfed maize, South Asia

Low input irrigated rice, South Asia

High input irrigated rice, South Asia

Page 204: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

204

Low input rainfed wheat, South Asia

High input rainfed wheat, South Asia

Low input irrigated wheat, South Asia

High input irrigated wheat, South Asia

Page 205: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

205

Figure 2.15 Yield effects from the neural network analysis

Low input rainfed wheat, Central Asia

Page 206: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

206

Figure 2.16. Total rice production by region in year 2000 and 2050 under scenarios with and without climate change.

Page 207: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

207

Figure 2.17. Total cereal demand by region in year 2000 and 2050 under scenarios with and without climate change.

Page 208: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

208

Figure 2.18. Total meat demand by region in year 2000 and 2050 under scenarios with and without climate change.

Page 209: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

209

Figure 2.19 Global commodity-level maize prices in year 2000 and 2050 under scenarios with and without climate change.

Page 210: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

210

Figure 2.20 Global commodity-level rice prices in year 2000 and 2050 under scenarios with and without climate change.

Page 211: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

211

Figure 2.21 Global commodity-level wheat prices in year 2000 and 2050 under scenarios with and without climate change.

Page 212: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

212

Figure 2.22 Global commodity-level soybean prices in year 2000 and 2050 under scenarios with and without climate change.

Page 213: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

213

Figure 2.23 Global commodity-level beef prices in year 2000 and 2050 under scenarios with and without climate change.

Page 214: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

214

Figure 2.24 Global commodity-level pork prices in year 2000 and 2050 under scenarios with and without climate change.

Page 215: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

215

Figure 2.25 Global commodity-level poultry prices in year 2000 and 2050 under scenarios with and without climate change.

Page 216: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

216

Figure 2.26 Aggregate cereal net trade by region in year 2000 and 2050 under scenarios with and without climate change.

Page 217: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

217

Figure 2.27 Aggregate meat net trade by region in year 2000 and 2050 under scenarios with and without climate change.

Page 218: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

218

Figure 2.28. Impact of climate change on agricultural GDP in Southeast Asia, 2080 (% changes from respective reference projection)

Source: Fischer, Shah, and Van Velthuizen 2002.

‐6

‐5

‐4

‐3

‐2

‐1

0

1

2

A1F1 A2 B2 B1 A2 B2 B1 A2 B2 A2 B2

HadCM3 CSIRO CGCM2 NCAR

Page 219: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

219

Figure 2.29 Global distribution of flood risk

Source: Dilley et al. 2005.

Page 220: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

220

Figure 2.30. Changes in global cereal prices under seven SRES scenarios with and without CO2 effects, relative to the reference scenario (no climate change)

Source: Parry et al. 2004.

Page 221: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

221

Figure 2.31. Estimated baseline population at risk in 1990 (A) and estimated population at risk in 2085 (B)

Note: Results of a logistic regression model with vapor pressure (humidity) as the predictor of dengue fever risk, using climate data from 1961 to 1990 (A). Forecast geographical distribution of dengue transmission based on climate projections for 2080–2100 from a global circulation model (CCGCMA2) (B). Colors represent probability of dengue fever transmission.

Source: Hales et al. 2002.

Page 222: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

222

Figure 2.32. Total number of malnourished Children in developing Asia: 2000 and 2050 (thousands of children, under 5 yrs of age)

Source: IMPACT model simulations

0

10000

20000

30000

40000

50000

60000

70000

80000

Central Asia East Asia South Asia Southeast Asia

Thousands of Child

ren

Year 2000 Baseline

Climate change

No Climate Change

Page 223: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

223

Figure 2.33. Total number of malnourished Children in China and India: 2000 and 2050 (thousands of children, under 5 years of age)

Source: IMPACT model simulations.

0

10000

20000

30000

40000

50000

60000

China India

Thousands of Child

ren

No ClimateChange

No ClimateChange

ClimateChange

ClimateChange

Year 2000 

Baseline

Year 2000

Baseline

Page 224: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

224

Figure 2.34. Intensity and causes of world environmental conflicts, 1980–2005

Source: Carius et al. 2006 in Schubert et al. 2008.

Page 225: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

225

Figure 3.1. Cumulative humanitarian risk hotspots

Note: This map shows cumulative humanitarian risk hotspots for three climate-related hazards—floods, cyclones, and droughts. Areas at risk for more than one type of hazard are considered to be of most concern for humanitarian actors. Yellow indicates one hazard, green two hazards, and blue all three hazards.

Source: Ehrhart et al. (2008).

Page 226: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

226

Figure 3.2. Geographical distribution of vulnerability in 2100 along an A2 emissions scenario with a climate sensitivity of 5.5°C

Source: Yohe (2006).

Page 227: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

227

Figure 3.3. Climate change vulnerability map of Southeast Asia

Note: For the legend, the scale used is 0-1 indicating the lowest vulnerability level (0) to the highest vulnerability level (1).

Source: Yusuf and Francisco (2009).

Page 228: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

228

Figure 3.4. Countries vulnerable to climate change

Source: Based on Table 3.3.

Page 229: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

229

Figure 4.1. Total technical mitigation potentials (all practices, all GHGs) for each region by 2030

Note: Boxes show one standard deviation above and below the mean estimate for per-area mitigation potential, and the bars show the 95% confidence interval about the mean. Based on the B2 scenario, although the pattern is similar for all SRES scenarios.

Source: Smith et al. (2007a), drawn from data in Smith et al. (2007b).

Page 230: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

230

Figure 4.2. Share of global GHG emissions by sector, 2000

Source: Drawn from data from WRI (2008).

Energy63%

Industrial Processes

3%

Agriculture13%

Land‐Use Change & Forestry18%

Waste3%

Page 231: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

231

Figure 4.3. Sources of emissions from the agricultural sector, 2000

Source: Drawn from data presented in (USEPA 2006).

Rice (CH4)11%

Residue Burning/Forest 

Clearning13%

Manure Management (CH4 & N2O)

7%

Livestock (CH4)32%

Fertilizers (N2O)37%

Page 232: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

232

Figure 4.4. Projected agricultural emissions by sector, 1990–2020

Source: Drawn from data used in USEPA (2006).

0

1,000

2,000

3,000

4,000

5,000

6,000

7,000

8,000

1990 2000 2010 2020

Emis

sion

s (M

tCO

2eq) CH4OtherAg

CH4ManureMgmt

CH4Rice

CH4EntericFerm

N20OtherAg

N20ManureMgmt

N20Soils

Page 233: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

233

Figure 4.5. CH4 and N2O emissions from historical and projected data in the agricultural sector, 1990–2020

Source: Drawn from data presented in (USEPA 2006).

Page 234: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

234

Figure 4.6. Global technical mitigation potential by 2030 of each agricultural management practice showing the impacts of each practice in GHG

Source: Smith et al. (2007a).

Page 235: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

235

Figure 4.7. Global estimates of economic mitigation potential for agricultural mitigation at different carbon prices at 2030

Notes: US-EPA (2006b) figures are for 2020 rather than 2030. Values for top-down models are taken from ranges given in Figure 8.7 of Smith et al. 2007a. Source: Smith et al. (2007a).

Page 236: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

236

Figure 4.8. Economic potential for GHG agricultural mitigation by 2030 at a range of prices of CO2-eq

Source: Smith et al. (2007b).

Page 237: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

237

Figure 5.1. Role of institutions in climate change responses.

Source: Meinzen-Dick and Moore (2009).

T i m eShort Long

Space

Plot

Com-munity

Nation

Global

Property Rights

Coordination

State

Col

lect

ive

Act

ion

ForestsReservoirs

Watershed management

PondsTerracing

Change variety

Carbon Markets

AgroforestrySoil Carbon

IPM

Irrigation

Seed Systems

International

New seed

Page 238: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

238

ANNEXES

Annex 1.1. List of ADB’s Developing Member Countries, by Region (as classified in the Annual Asian Development Outlook Publication)

Region Countries

A. Central Asia 1. Armenia 2. Azerbaijan 3. Georgia 4. Kazakhstan 5. Kyrgyz Republic 6. Tajikistan 7. Turkmenistan 8. Uzbekistan B. East Asia 9. China, People’s Republic of 10. Hong Kong, China 11. Korea, Republic of 12. Mongolia 13. Taipei,China C. South Asia 14. Afghanistan, Islamic Republic of 15. Bangladesh 16. Bhutan 17. India 18. Maldives 19. Nepal 20. Pakistan 21. Sri Lanka D. Southeast Asia 22. Cambodia 23. Indonesia 24. Lao People’s Democratic Republic 25. Malaysia 26. Myanmar 27. Philippines 28. Singapore 29. Thailand 30. Viet Nam E. The Pacific 31. Cook Islands 32. Fiji Islands 33. Kiribati 34. Marshall Islands, Republic of 35. Micronesia, Federated States of 36. Nauru 37. Palau, Republic of 38. Papua New Guinea 39. Samoa 40. Solomon Islands 41. Timor-Leste, Democratic Republic of 42. Tonga 43. Tuvalu 44. Vanuatu Total DMCs 44

Page 239: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

239

List of Subregions in Asia and the Pacific

Subregions Member Countries

1. Greater Mekong Subregion (GMS) 1. Cambodia 2. Lao People's Democratic Republic 3. Myanmar 4. China, People’s Republic of 5. Thailand 6. Viet Nam

2. Central Asia Regional Economic Cooperation (CAREC)

1. Afghanistan, Islamic Republic of

2. Azerbaijan 3. Kazakhstan 4. Kyrgyz Republic 5. Mongolia 6. China, People’s Republic of 7. Tajikistan 8. Uzbekistan

3. South Asia Subregional Economic Cooperation (SASEC)

1. Bangladesh

2. Bhutan 3. India 4. Nepal

4. Indonesia-Malaysia-Thailand Growth Triangle (IMT-GT)

1. Indonesia

2. Malaysia 3. Thailand 5. Brunei Darussalam-Indonesia-Malaysia-Philippines East ASEAN Growth Area (BIMP-EAGA)

1. Brunei Darussalam

2. Indonesia 3. Malaysia 4. Philippines 6. South Asian Association for Regional Cooperation (SAARC)

1. Bangladesh

2. Bhutan 3. India 4. Maldives 5. Nepal 6. Pakistan 7. Sri Lanka 7. Subregional Economic Cooperation in South and Central Asia (SECSCA)

1. Afghanistan, Islamic Republic of

2. Pakistan 3. Tajikistan 4. Turkmenistan 5. Uzbekistan

8. Shanghai Cooperation Organization (SCO) 1. Kazakhstan

Page 240: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

240

Subregions Member Countries

2. Kyrgyz Republic 3. China, People’s Republic of 4. Tajikistan 5. Uzbekistan

9. South Asia Growth Quadrangle (SAGQ) 1. Bangladesh 2. Bhutan 3. India 4. Nepal

10. Bay of Bengal Initiative for Multi-Sectoral Technical and Economic Cooperation (BIMSTEC)

1. Bangladesh

2. Bhutan 3. India 4. Myanmar 5. Nepal 6. Sri Lanka 7. Thailand

11. Pacific Plan 1. Cook islands 2. Fiji Islands 3. Kiribati 4. Marshall Islands, Republic of 5. Micronesia, Federated States of 6. Nauru 7. Palau 8. Papua New Guinea 9. Samoa 10. Solomon Islands 11. Tonga 12. Tuvalu 13. Vanuatu 14. Australia 15. New Zealand 12. Asia Cooperation Dialogue (ACD) 1. Bangladesh 2. Bhutan 3. Cambodia 4. India 5. Indonesia 6. Kazakhstan 7. Lao People’s Democratic Republic 8. Malaysia 9. Mongolia 10. Myanmar 11. Pakistan 12. China, People’s Republic of 13. Philippines 14. Korea, Republic of 15. Singapore 16. Sri Lanka 17. Thailand 18. Viet Nam 19. Brunei Darussalam

Page 241: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

241

Subregions Member Countries

20. Japan 13. Ayeyawady-Chao Phraya-Mekong Economic Cooperation Strategy (ACMES)

1. Cambodia

2. Lao People’s Democratic Republic 3. Myanmar 4. Thailand 5. Viet Nam 14. Asia-Pacific Economic Cooperation (APEC) 1. Hongkong, China 2. Indonesia 3. Malaysia 4. Papua New Guinea 5. China, People’s Republic of 6. Philippines 7. Korea, Republic of 8. Singapore 9. Taipei,China 10. Thailand 11. Viet Nam 12. Australia 13. Brunei Darussalam 14. Japan 15. New Zealand 15. Association of Southeast Asian Nations (ASEAN)

1. Cambodia

2. Indonesia 3. Lao People’s Democratic Republic 4. Malaysia 5. Myanmar 6. Philippines 7. Singapore 8. Thailand 9. Viet Nam 10. Brunei Darussalam 16. ASEAN plus People’s Republic of China, Japan, and Republic of Korea (ASEAN + 3)

1. Cambodia

2. China, People’s Republic of 3. Indonesia 4. Korea, Republic of 5. Lao People’s Democratic Republic 6. Malaysia 7. Myanmar 8. Philippines 9. Singapore 10. Thailand 11. Viet Nam 12. Brunei Darussalam 13. Japan

17. Asia–Europe Meeting (ASEM) 1. Cambodia

Page 242: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

242

Subregions Member Countries

2. Indonesia 3. Lao People’s Democratic Republic 4. Malaysia 5. Myanmar 6. China, People’s Republic of 7. Philippines 8. Korea, Republic of 9. Singapore 10. Thailand 11. Viet Nam 12. Brunei Darussalam 13. Japan

Page 243: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

243

Annex 2.1. Models Used for Agricultural Climate Change Impact Analysis.

In 1996, a study highlighted that despite the many studies on global warming, there is no consensus on the impacts of three major variables that may mitigate the impact of climate change on agriculture: the magnitude of regional changes in temperature and precipitation, the magnitude of the beneficial effects of higher CO2 on crop yields, and the ability of farmers to adapt to climate changes (Wolfe 1996). More than 10 years later, such observation remains valid, resulting in major disagreements among studies about the impacts of climate change on agriculture in all regions of the world, including the Asia and Pacific region.

Models assessing the impacts of climate change on agriculture have been placed into three categories: agronomic-economic simulations, agroecological zone analysis, and Ricardian cross-sectional analyses (Mendelsohn and Dinar 1999). The agronomic-economic simulations use crop models that contain data from carefully controlled experiments, which vary climate and carbon dioxide levels to simulate different climate change scenarios. Economic impacts are then estimated by inputting yield results from the experiments into economic models. Agroecological zone analyses assign crops to specific agroecological zones, and then determine expected yields as well as the impacts that climate change will have on those yields. Ricardian models measure the economic performance of farms in different climatic regions to determine the effect of changes in climate on this performance. In addition to the basic differences due to the structure of these models, additional variance in results can be observed due to the extent of farmer adaptation and CO2 fertilization impact on crop yields included in the models. These aspects are discussed further in subsequent sections.

Agronomic-Economic Simulations

Agronomic-economic models consist of a crop model that uses output results of experiments controlled for climate and CO2 concentrations. Crop model results are then inputted into an economic model to determine crop prices, outputs and net revenues. Mendelsohn and Dinar (1999) point out that most of these models generally focus on a small selection of crops (generally grains) since expansive experimentation is required for each of the crops to be included in the model.

Several studies mentioned in chapter 2, including, Parry et al. 1999; Parry et al. 2004; and Lin et al. 2005, use crop models to assess the impact of global warming on yields. Parry et al. (1999) and Parry et al. (2004) use the IBSNAT-ICASA model family to estimate yield responses to temperature and CO2 level, including CERES-Wheat, CERES-Rice, CERES-Maize, and SOYGRO (for soybean).

In Parry et al. (1999), simulations were specified and validated in 124 sites in 18 countries under a number of climate change scenarios. Those simulations were then aggregated to agroclimatic regions to statistically derive regional yield response functions for using in an integrated assessment model. These functions took the form of multiple linear and quadratic regression models to reflect the combined changes in temperature, precipitation and CO2 concentration (Zhu 2007).

The same approach was followed by Parry et al. (2004) that evaluated a broader range of climate change scenarios. According to the authors, projected changes in yield were calculated using transfer functions derived from crop model simulations with observed climate data and

Page 244: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

244

projected climate change scenarios. The impacts of climate change were estimated for scenarios developed from the HadCM3 global climate model under the IPCC SRES A1F1, A2, B1, and B2 scenarios. The authors used production functions that incorporate: i) crop responses to changes in temperature and precipitation with the current management; ii) crop responses to temperature and precipitation with farm-level and regional adjustments; and iii) crop responses to carbon dioxide. Yield responses to combined changes in temperature and precipitation were taken from over 50 previously published and unpublished regional climate impact studies. Farm-level adaptation strategies included changes in planting date and application of additional fertilization and irrigation in the current irrigated areas. Finally, the basic linked system (BLS) was used to evaluate consequent changes in global cereal production, cereal prices and the number of people at risk from hunger. The impacts of climate change on arable land were not considered although the upper limit of available arable land (based on historical climate conditions) for expansion of crop production was considered based on FAO database (Zhu 2007).

According to Zhu (2007), the advantages of crop simulation models over statistical models are that they can simultaneously consider multiple factors that affect crop growth and that the models are based on the physiological process of crop growth. Process-based crop models are more robust for extrapolation than purely statistically-based models. When statistical models are applied to a different environmental, the parameters need to be re-estimated (calibrated), and usually there are more concerns about the model structure or the functional forms of model equations. However, crop models also need to be calibrated against experimental data, which might be a problem when applied to significantly different environmental conditions, particularly in global assessments (Zhu 2007). A criticism of the agronomic studies is that they fail to account for adaptations that farmers continuously undertake and therefore possibly overestimating negative impacts of climate change (Kurukulasuriya and Ajwad 2007).

Agroecological Zone Analysis

Agro-ecological zone (AEZ) models assign particular crops to certain agroecological zones, and then estimate yields for the different zones. The model reacts to changes in climate by altering both the agroecological zones and the crops being produced in the zones. In this way, the models can estimate the impact of climate change on crop yields. As in the agronomic-economic models, these results can then be applied to economic models to determine any supply or market impacts (Mendelsohn and Dinar 1999). According to Fischer et al. (2002), the AEZ methodology provides a standardized framework for the characterization of climate, soil and terrain conditions relevant to agricultural production. Crop modeling and environmental matching procedures are used to identify crop-specific limitations of prevailing climate, soil and terrain resources, under assumed levels of inputs and management conditions (Fischer et al. 2002). As a result, maximum potential and agronomically attainable (potential) crop yields for basic land resources units are provided.

The Global AEZ project and associated climate change studies done by the Food and Agricultural Organization of the United Nations (FAO) and the International Institute for Applied System Analysis (IIASA) provide a comprehensive assessment of climate change impacts on crop areas at global scale (see Fischer et al. 2001, 2002) (Zhu, 2007).

An agroecological zone study indicates that the magnitude of temperature increase and change in rainfall amount will affect the projected area suitable for cereal production (Fischer et al. 2001). At a global level, the amount of cultivable land was found to increase with a 2ºC

Page 245: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

245

increase in temperature and no change in rainfall amounts. A 3ºC increase and no change in rainfall amounts, however, led to a decline in amount of cultivable rainfed land from the 2ºC increase scenario. Additional results showed that with adaptation of crop calendars, switching crop types, and yield increases due to CO2 fertilization effect incorporated into the model, climate change was found to benefit developed countries more than developing countries when allowing for one rainfed crop per year, multi-crop rainfed production or irrigated production.

Ricardian Models

Ricardian models use a cross-sectional approach to analyze the impacts of climate change and other factors on land values and farm revenues (Mendelsohn, Nordhaus and Shaw 1994). This type of model differs from the two discussed previously in that it incorporates farmers’ ability to adjust the inputs or technology used to adapt to a warmer climate into the model. Hence, results from Ricardian models have generally shown a more positive outlook for future agricultural production than the agronomic-economic and agroecological models. The inability to control the experiments across farms is one disadvantage to this type of model compared to the other model types (Mendelsohn and Dinar 1999). Some other criticisms that have been raised against the Ricardian method is that it might overestimate benefits and that it uses constant output and other input prices (Kurukulasuriya and Ajwad 2007).

As presented in Chapter 2, Seo et al. (2005) and Kurukulasuriya and Ajwad (2007) use the Ricardian method to estimate the impacts of climate change on agricultural net revenue in Sri Lankan in different climate zones. The model captures adaptation implicitly by comparing net outcomes for farmers facing different zones. It is assumed that farmers maximize net revenues per hectare. Therefore, given household preferences and endowments, farmers will choose the best adaptation strategy available. The Ricardian model regresses net revenue on climate and other explanatory variables. This analysis – cross-sectional observation across different climates – can then reveal climate sensitivity of farms. In both studies, several GCM scenarios were used to project impacts of climate change on agricultural income.

Cline (2007) assesses impacts of climate change on agriculture at national or sub-national level based on two models framework: Ricardian models and crop models. Cline’s study was based on the idea of model averaging and ensemble forecasting. In ensemble forecasting or model averaging, instead of selecting a “best” model, the modeler can combine the predictions of different candidate models to get the more robust prediction. This practice has become popular in recent years and is considered a promising method to deal with the uncertainty in the specifications of model structure. The author arrived at preferred estimates by synthesizing those two sets of estimates and using them as a basis for new estimates. The first - the Ricardian cross-section models - relates agricultural capacity statistically to temperature and precipitation on the basis of statistical estimates from farm survey or county-level data across climatic zones (Cline 2007). Studies are available for the United States, Canada, many countries in Africa, countries in Latin America and India17. The author applies these country-specific models to estimate impacts in countries accounting for 35 percent of global agricultural output and about half of the number of countries. In countries where such studies are not available, the author applied the Mendelsohn-Schlesinger Ricardian model for the United States for climate estimates. In these

17 Mendelsohn and Schlesinger, 1999 (United States), Reinsborough, 2003 (Canada), Kurukulasuriya, 2006 (Africa), World

Bank farm surveys (Latin America) and Mendelsohn, Dinar, and Sanghi, 2001 (India).

Page 246: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

246

cases, however, the weighting given to Ricardian estimates is reduced and weighting of crop models is increased.

The second framework is based on crop models and consists of region-specific calculations synthesized from estimates by agricultural scientists in 18 countries as applied to alternative GCM projections of climate scenarios. Crop model estimates are from Rosenzweig and Iglesias (2006) (Cline 2007).

Farmer Adaptation

Adaptation to climate change by farmers has been shown to decrease the overall impacts of climate change in many models. Darwin et al. (1995), for example, found that although yield losses were 20 to 30 percent without adaptation, adaptation led to slightly increased yields at the global level. While many models include the potential for farmer adaptation, there is considerable disagreement as to the extent and costs of that adjustment.

Mendelsohn and Dinar (1999) suggest that many agronomic models (both agronomic-economic and agroecological) may present an overly pessimistic view of climate change impacts because they either do not include farmer adaptation at all, or do not include it to the extent that farmers are likely to adapt in reality. Other authors disagree, stating that adaptation costs could be larger than those presented in the Ricardian models, thus hindering the ability of farmers in developing countries to change their practices (Quiggin and Horowitz 1999). Quiggin and Horowitz (1999) suggest that the Ricardian approach is biased by assuming infinite adjustment costs, and would yield the lower-bound estimate of the costs of climate change. Changes in climate could lead to adaptations that would require quite expensive capital costs, specifically those for irrigation. In addition to the capital costs that may be incurred in some areas to build new irrigation systems, changes in climate may render existing systems useless, thus leading to the scrapping of long-lived capital costs in these areas (Quiggin and Horowitz 1999). Reilly (1999b) also mentions the potential for capital expenditure losses if farming equipment is rendered useless due to a change in crop planted.

In addition to these increases in capital costs, other adaptation costs may also be incurred, such as migration and labor adjustment costs, particularly in developing countries. Rosenzweig (2000) suggests that adaptation may be slower in developing countries due to the absence of infrastructure for agricultural research that exists in developed countries. Reilly (1999a) points to prior research that suggests grain yields in developing countries could be lower with adaptation than without due to the low prices caused by greater adaptation in the developed countries. Other authors mention potential water availability and environmental costs caused by converting ecologically sensitive land into cropland to adjust to new regional climate patterns. Tropical regions may suffer negative impacts from droughts, due to the nonlinear relationship between temperature and evapotranspiration, even though climate changes in these regions are expected to be less; these regions will also face greater difficulties in shifting planting dates, as they are limited more by rainfall than temperature (Reilly 1995). Potential species and ecosystem impacts from shifts in cropped areas, and increased stresses on water availability have been mentioned as possible environmental costs (Lewandrowski and Schimmelpfennig 1999).

Page 247: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

247

References

Adams, R. M. and B. H. Hurd. 1999. Climate change and agriculture: Some regional implications. Choices 14(1): 22-23.

Cline, W. 2007. Global Warming and Agriculture: Impact Estimates by Country. Washington, DC: Peterson Institute.

Darwin, R., M. Tsigas, J. Lewandrowski, and A. Raneses. 1995. World agriculture and climate change. Economic Adaptations. Agricultural Economic Report Number 703, Washington, D.C.: United States Department of Agriculture.

Downing, T. E. 1993. The effects of climate change on agriculture and food security. Renewable Energy Vol. 3 (4/5): 491-97.

Fischer, G., M. Shah, H. van Velthuizen, and F.O. Nachtergaele. 2001. Global agro-ecological assessment for agriculture in the 21st century. Vienna, Austria: International Institute for Applied Systems Analysis.

Fischer, G., M., Shah, and H. van Velthuizen.2002. Climate change and agricultural vulnerability, a special report prepared by the International Institute for Applied Systems Analysis under United Nations Institutional Contract Agreement No. 1113 on “Climate Change and Agricultural Vulnerability” as a contribute to the World Summit on Sustainable Development, Johannesburg 2002.

Lin, E., W. Xiong, H. Ju, Y. Xu, Y. Li, L. Bai and L. Xie. 2005. Climate change impacts on crop yield and quality with CO2 fertilization in China, Philosophical Transaction of the Royal Society (B), 360: 2149-2154.

Kurukulasuriya, P., and M. I. Ajwad. 2007. Application of the Ricardian technique to estimate the impact of climate change on smallholder farming in Sri Lanka. Climatic Change 81 (1): 39-59

Mendelsohn, R., and A. Dinar. 1999. Climate change, agriculture, and developing countries: Does adaptation matter? The World Bank Research Observer 14(2): 277-93.

Parry, M., C. Rosenzweig, A. Iglesias, G. Fischer, and M. Livermore. 1999.Climate change and world food security: a new assessment. Global Environmental Change 9: S51-S67.

Parry, M.L., C. Rosenzweig, A. Iglesias, M. Livermore and G. Fischer 2004. Effects of climate change on global food production under SRES emissions and social-economic scenarios. Global Environmental Change 14: 53-67.

Quiggin, J. and J. K. Horowitz. 1999. The impact of global warming on agriculture: A Ricardian analysis: A comment. American Economic Review 89 (4): 1044-1045.

Reilly, J. 1999a. What does climate change mean for agriculture in developing countries? A comment on Mendelsohn and Dinar. The World Bank Research Observer 14 (2): 295-305.

Page 248: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

248

Reilly, J. 1999b. Climate change: Can agriculture adapt? Choices 14 (1): 4-8.

Reilly, J. 1995. Climate change and agriculture - Research findings and policy considerations. In Population and food in the early twenty-first century: Meeting future food demand of an increasing population, ed. Nurul Islam, Washington, D.C.: IFPRI.

Rosenzweig, C., and A. Iglesias. 2006. Potential Impacts of Climate Change on World Food Supply: Data Sets from a Major Crop Modeling Study. New York: Goddard Institute for Space Studies, Columbia University. Available at http://sedac.ciesin.columbia.edu (accessed August 9, 2006).

Wolfe, D. 1996. Potential impact of climate change on agriculture and food supply. In Sustainable development and global climate change: Conflicts and connections, eds. James White, William R. Wagner, and Wendy H. Petry. Proceedings of a conference sponsored by the Center for Environmental Information, Inc., 4-5 Dec 1995, in Washington, D.C., and published with the assistance of the US Global Change Research Program (USGCRP).

Zhu, T. 2007. Review of the Impacts of Future CO2 and Climate on Agricultural Production.

Page 249: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

249

Annex 2.2. IFPRI’s Climate Change Modeling Framework

Approach

IFPRI implemented this research through an intensive desk study, compiling and critically synthesizing and analyzing existing analyses, secondary data, case studies, and information on climate change, agriculture, and other relevant literature from a large variety of sources. Quantitative analyses supplemented the desk study by applying a modeling system linking several models that provide scenarios of the important impacts of climate change on agriculture to 2050 and a methodology for assessing adaptation costs. The modeling results are compared with the results from the comprehensive synthesis of the existing climate change impact models.

Modeling overview

General equilibrium models generally divide the world into 15 to 30 regions with very limited disaggregation at the country or within country scale. Partial equilibrium model generally have a greater detail on the sector—here agriculture—but rely on economic relationships neglecting some or all local biophysical settings. However, in the real world field-level production decisions made by farmers are influenced by variables that include relatively unchanging geophysical variables such as elevation, slope, and soil characteristics, climate variables of precipitation, temperature and available solar radiation, and economic variables such as prices, property rights, and social infrastructure.

The modeling framework used here reconciles the often limited resolution of macro-level economic models that operate through equilibrium-driven relationships at a national or even more aggregate regional level with detailed models of dynamic biophysical processes. In particular, we link crop growth model results with a neutral-network to allocate results across landscapes. These results are then fed into a partial agricultural equilibrium model. Linking these types of models is needed to assess the impacts of climate change and the potential for climate change mitigation and adaptation policies and programs.

An illustrative schematic of the linkage between the global agricultural policy and trade modeling of the partial agriculture equilibrium model with the agronomic potential modeling is shown in Figure 1. We see that the main climate change effects occur on the production side while most of the key welfare implications are derived from the demand side results.

Modeling climate change impacts on agriculture

The challenge of modeling climate change impacts arises in the wide ranging nature of characteristics and processes that underlie the working of markets, ecosystems, and human behavior. Our analytical framework integrates modeling components that range from the macro to the micro and from processes that are driven by economics to those that are essentially biological in nature. Considering this entire range provides a more holistic assessment of the consequences of climate change and the benefits that can be generated by well-designed climate change mitigation and adaptation policies and programs. Simulation techniques that integrate physical and economic models are used to investigate the effects on rural producers under a range of climate and socioeconomic futures.

The climate change modeling system combines a biophysical model (the DSSAT crop modeling suite) of responses of selected crops to climate, soil and nutrients with the ISPAM data set of crop location and management techniques (You and Wood 2006) and IFPRI’s global

Page 250: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

250

agricultural supply and demand projections model, IMPACT18. IMPACT’s detailed partial-equilibrium representation of agricultural production and consumption is enhanced by a detailed biophysical representation of the response of key crops to climate and nutrient changes. This modeling framework is used to undertake economic and policy scenario analysis of the impacts that new crop technologies and improved management can have on agricultural economies, as well as the impact of crop prices, fertilizer prices, investments in irrigation, and fertilizer on agricultural productivity. Summary descriptions of the models utilized in the linked system are provided below.

Adaptation needs and potential

Climate change will bring location-specific changes in precipitation and variability and temperature levels and variability. Ongoing research at IFPRI and other institutions has identified the agro-climatic suitability for the world’s crops globally, given today’s climate.

This research uses the range of climate conditions expected in 2050 to assess how suitability would change. The location-specific change in suitability for existing crops provides a clear indication of where adaptation efforts would need to be focused. These results are relevant to a variety of audiences—from ADB’s decision-makers and macroeconomic and trade policymakers, for whom the need to rely increasingly on staple imports would be of high interest; to infrastructure planners, for whom the location of newly important agricultural areas should influence road, rail, and irrigation investments; to agricultural research managers, for whom the extent of new suitability environments should be an important factor in research investments.

Model descriptions

The modeling environment consists of three distinct software models and related databases – the DSSAT crop model, a neural net representation of crop model climate interactions, and IFPRI’s IMPACT2009 model.

Crop modeling

The DSSAT crop simulation model is an extremely detailed process model of the daily development of a crop from planting to harvest-ready. It requires daily weather data, including maximum and minimum temperature, solar radiation, and precipitation, a description of the soil physical and chemical characteristics of the field, and crop management, including crop, variety, planting date, plant spacing, and inputs such as fertilizer and irrigation. For maize, wheat, rice, groundnuts, and soybeans, we use the DSSAT crop model, version 4.0 (Jones et al. 2003).

For mapping these results to other crops in IMPACT the primary assumption is that plants with similar biophysical metabolic pathways will react similarly to any given climate change effect in a particular geographic region. IMPACT crops are classed as using either the C3 or C4 photosynthetic pathways to produce the basic carbohydrates that make up the plant material. Millet, sorghum, sugarcane, and cotton all use the C4 pathway and are assumed to follow the DSSAT results for maize, which is also C4, in the respective geographic regions. The remainder of the crops all follow the C3 pathway and the average climate effects from wheat,

18 IMPACT - International Model for Policy Analysis of Agricultural Commodities and Trade. See Rosegrant et al. 2008 for

details at http://www.ifpri.org/themes/impact/impactwater.pdf

Page 251: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

251

rice, soy, and groundnut from the same geographic region are used, with two exceptions. The IMPACT commodities of “other grains” and dryland legumes are directly mapped to the DSSAT results for wheat and groundnut, respectively.

Climate data

To simulate today’s climate we use the Worldclim current conditions data set (www.worldclim.org) which is representative of 1950-2000 and reports monthly average minimum and maximum temperatures and monthly average precipitation. Site-specific daily weather data are generated stochastically using the SIMMETEO software.

Precipitation rates and solar radiation data were obtained from NASA’s LDAS website (http://ldas.gsfc.nasa.gov/). We used the results from the Variable Infiltration Capacity (VIC) land surface model. For shortwave radiation, monthly averages at 10 arc-minute resolution were obtained for the years 1979-2000. Overall averages for each month were computed between all the years (e.g., the January average was computed as [January 1979 + January 1980 + ... + January 2000 ] / 22).

Rainfall rates were obtained at three-hourly intervals for the years 1981, 1985, 1991, and 1995. A day was determined to have experienced a precipitation event if the average rainfall rate for the day exceeded a small threshold. The number of days experiencing a rainfall event within each month was then counted up and averaged over the four years.

The monthly values were regressed nonlinearly using the Worldclim monthly temperature and climate data, elevation from the GLOBE dataset (http://www.ngdc.noaa.gov/mgg/topo/globe.html) and latitude. These regressions were used to estimate monthly solar radiation data and the number of rainy days for both today and the future. These projections were then used by SIMETEO to generate the daily values used in DSSAT.

For future climate in this report, we use the Worldclim future climate data set for 2050 downscaled to 5 arc minutes from the AR3 Hadley GCM run with the A2a forcings scenario (http://www.worldclim.org/futdown.htm). At one time this was considered an extreme scenario although recent findings suggest it may not be. We assume that all climate variables change linearly between their values in 2000 and 2050. This assumption eliminates any random extreme events such as droughts or high rainfall periods and also assumes that the forcing effects of GHG emissions proceeds linearly, that is, we do not see a gradual speedup in climate change.

A brief description and characterization of the “family” of scenarios used in the 3rd and 4th IPCC assessments is shown in Figure 2, to give the reader a better idea of the assumptions on underlying driving forces of change.

Other agronomic inputs

Six other agronomic inputs are key – soil characteristics, crop variety, cropping calendar, CO2 fertilization effects, irrigation and nutrient levels.

Soil characteristics

The DSSAT model uses many different soil characteristics in determining crop progress through the growing season. John Dimes of ICRISAT and Jawoo Koo of IFPRI collaborated to classify the FAO soil types into 27 meta-soil types. Each soil type is defined by a triple of soil organic

Page 252: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

252

carbon content (high/medium/low), soil rooting depth as a proxy for available water content (deep/medium/shallow), and major constituent (sand/loam/clay).

Crop variety

DSSAT includes many different varieties of each crop. For the results reported here, we used the maize variety Garst 8808, a USA winter wheat variety, a Large-seeded Virginia runner type groundnut variety and an IRRI recent rice variety.

Cropping calendar

Climate change will alter the cropping calendar. In some locations crops can be grown in 2000 but not in 2050, or vice versa. This has been taken into account.

CO2 fertilization effects

DSSAT has an option to include CO2 fertilization effects. We simulate a range of levels of atmospheric CO2 from 300 to 900 ppm.

Irrigation

We consider two types of water availability. Rainfed crops receive water either from precipitation at the time it falls or from soil moisture. Soil characteristics influence the extent to which previous precipitation events provide water for growth in future periods. Irrigated crops receive water automatically in the DSSAT model whenever soil moisture falls below 90 percent of the holding capacity.

Nutrient level

The DSSAT model only incorporates application of nitrogen exogenously. We vary the amount of elemental N from 0 to 300 kg per hectare. For low-

From DSSAT to a reduced form estimating function – the CM-NN output

The DSSAT crop model (CM) is computationally intense. To allow multiple simulations of climate effects for the entire surface of the globe, we developed a reduced form implementation. We ran the crop model for each crop and variety with a wide range of climate and agronomic inputs and then estimated a feed-forward neural net (NN) for each of the 27 soil categories. We obtain a continuous and differentiable approximation of the crop model results that allows us to find the maximum possible yield and corresponding nitrogen input needed based on location-specific geophysical characteristics and climate. The results of this estimation process are fed into the IMPACT model.

The IMPACT Model19

The IMPACT model was developed by the International Food Policy Research Institute (IFPRI) for projecting global food supply, food demand and food security to year 2020 and beyond (Rosegrant et al. 2001). It is a partial equilibrium agricultural model for crop and livestock commodities, including cereals, soybeans, roots and tubers, meats, milk, eggs, oilseeds,

19 We provide an overview of the IMPACT model here and refer interested readers to Rosegrant et al. 2008 for technical

details.

Page 253: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

253

oilcakes/meals, sugar/sweeteners, and fruits and vegetables. It is specified as a set of 115 country and regional sub-models, within each of which supply, demand, and prices for agricultural commodities are determined. Large countries are further divided into major river basins. The result is 281 spatial units, called food production units (FPUs). The model links the various countries and regions through international trade using a series of linear and nonlinear equations to approximate the underlying production and demand functions. World agricultural commodity prices are determined annually at levels that clear international markets. Growth in crop production in each country is determined by crop and input prices, the rate of productivity growth, investment in irrigation, and water availability. Demand is a function of prices, income, and population growth and contains four categories of commodity demand—food, feed, biofuels feedstock, and other uses.

Modeling Climate Change in IMPACT

Climate change effects on crop productivity enter into the IMPACT model by affecting both crop area and yield. Yields are altered through the intrinsic yield growth coefficient, gCY, in the yield equation as well as the water availability coefficient (WAT) for irrigated crops.

);()1()()( tnitnitnitnkk

tnitnitni WATYCgCYPFPSYC ikniin (1)

We generate relative climate change productivity effects by calculating location specific yields for each of the five crops modeled with DSSAT for 2000 to 2050 and then constructing a ratio. The ratio is then used to increase or decrease gCY. Climate effects on water availability are captured as explained in Rosegrant et al (2008) for irrigated crops while rainfed crops react to changes in precipitation as modeled in the biophysical crop model DSSAT.

FPUs are large areas. For example, the Ganges FPU is the entire length of the Ganges River. Within an FPU, there can be large variation in climate and agronomic characteristics. A major challenge was to come up with an aggregation scheme to take outputs from the CM-NN functions to the IMPACT FPUs. The process we used proceeds as follows. First, within an FPU, use the ISPAM data set, with a spatial resolution of 5 arc-minutes (approximately 10 km at the equator), and estimate climate change yield productivity ratio for each crop/management combination within an ISPAM pixel. We then weight these ratios by the share of the crop in the FPU area and use the weighted ratio in the yield equation (1).

Harvested areas in the IMPACT model are affected by climate change in a similar way to yields, though with a slight complication. In any particular FPU, land may become more or less suitable for any crop and will impact the intrinsic area growth rate, gA, in the area growth calculation. Water availability will affect the WAT factor for irrigated and rainfed crops as with the yields.

);()1()()( tnitnitnitnjij

tnitnitni WATACgAPSPSAC ijniin

(2)

Crop calendar changes due to climate change caused two distinct issues. When the crop calendar in an area changes so that a crop that was grown in 2000 can no longer be grown in 2050, we implement an adjustment to gA that will bring the harvested area to zero—or nearly so—by 2050. However, when it becomes possible to grow a crop in 2050 where it could not be grown in 2000, we do not add this new area. An example is that parts of Ontario, Canada will be

Page 254: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

254

able to grow maize in 2050 that have too short a growing season now, according to the models used in this exercise. As a result our estimates of future production are biased downward somewhat. The effect is likely to be small, however, as new areas have other constraints on crop productivity, in particular soil characteristics.

References

Jones, J. W., et al. 2003. The DSSAT cropping system model. European Journal of Agronomy 18 (3-4): 235-265.

Rosegrant, M.W., S. Msangi, C. Ringler, T.B. Sulser, T, Zhu, and S.A Cline. 2008. International Model for Policy Analysis of Agricultural Commodities and Trade (IMPACT): Model Description. International Food Policy Research Institute: Washington, D.C. <ihttp://www.ifpri.org/themes/impact/impactwater.pdf

You, L. and S. Wood. 2006. An entropy approach to spatial disaggregation of agricultural production. Agricultural Systems 90 (1-3): 329-347.

Page 255: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

255

Figure 1. Linkage between land use and agricultural market models.

Ag Policy & Trade Model

Welfare Analysis

Yield Area

Consumption (+per cap calorie availability) )

Equilibrium Trade Balance

cropping calendar

Biofuel Feedstock Demand

Food/Feed

Price conversion technology

Production

DemandEnergy & Biofuels demand

Other Demand

GDP Popn (exog)

Agronomic Modeling

Nutritional status

Other key determinants: access to clean water female schooling

Soil characteristics

Crop yield potential

model “corrections

Ag R&D investments + i i i

Adaptations: ag technology & crop varieties

Climate scenarios: Temp Precipitation Atmos CO2,

Page 256: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

256

Growth-focused policy objectives Eco-friendly policies

More globally integrated

A1 More integrated world with

cooperation Rapid economic growth Global population reaches 9

billion by 2050 then declines gradually afterwards

Quick spread of new efficient technologies

B1 More integrated world with policies

more friendly towards environment and emphasis on global solutions to economic, social and environmental issues.

Rapid economic growth (like A1) – but more oriented towards a service-oriented information economy

Global population reaches 9 billion by 2050 then declines (like A1)

Reduction in materially-intensive consumption and introduction of clean/resource-efficient technologies

More divided geo-politically

A2 More divided world with less

cooperation b/w nations Regionally-oriented economic

development, with lower per capita growth

Continually-increasing population growth

Slower and more fragmented spread of technologies

B2 More divided, but still eco-friendly

world Intermediate levels of economic

development and growth Continually-increasing population (but

slower than under A2) Less rapid and more fragmented pattern

of technological change (compared to A1 and B1)

Figure 2. Characterization of global IPCC scenarios (SRES20).

20 SRES = Special Report on Emissions Scenarios. These scenarios were developed for the 3rd IPCC Assessment Report in

2001 and also used for the 4th (AR4) assessment in 2007, to make different assumptions for future greenhouse gas pollution, land use changes, and their underlying driving forces.

Page 257: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

257

Annex 3.1 Parameters for agricultural employment as share of total employment (reflecting sensitivity to climate change)

Country Agricultural employment as % of total

employment

High sensitivity Bhutan 93.6Nepal 93.0Timor-Leste 81.1

Lao People's Democratic Republic 75.8Papua New Guinea 72.0Solomon Islands 71.7Myanmar 68.9Cambodia 68.6Viet Nam 65.7Afghanistan 65.7China 64.4India 57.8Thailand 53.3Bangladesh 51.8Indonesia 45.7Pakistan 45.0Sri Lanka 44.3

Medium sensitivity Fiji 38.1Philippines 37.1Vanuatu 34.4Nauru 33.3Turkmenistan 31.9Samoa 31.8Tonga 31.6Tajikistan 31.2Cook Islands 28.6Azerbaijan 25.1Uzbekistan 25.0Marshall Islands 25.0Tuvalu 25.0Micronesia, Federated States of 24.4

Page 258: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

258

Kiribati 23.9Kyrgyzstan 23.4Palau 22.2Mongolia 21.5Maldives 19.3Georgia 17.8Kazakhstan 16.1Malaysia 15.9

Low sensitivity Armenia 10.9Korea, Republic of 7.7Singapore 0.1

Source: FAOSTAT (FAO 2004).

Page 259: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

259

Annex 3.2. Poverty incidence reflecting relative adaptive capacity in the Asia and Pacific region.

Country Poverty Incidence

PPP 2005

Low Adaptive Capacity Nepal 54.7 Bangladesh 50.47 Timor-Leste 43.56 India 41.64 Cambodia 40.19 Uzbekistan 38.81 Afghanistan Myanmar Lao PDR 35.68

Medium Adaptive Capacity Papua New Guinea 29.7 Bhutan 26.79 Viet Nam 22.81 Philippines 22.62 Pakistan 22.59 Mongolia 22.38 Kyrgyz Republic 21.81 Tajikistan 21.49 Indonesia 21.44 China 15.92 Georgia 13.44 Turkmenistan 11.72 Sri Lanka 10.33

High Adaptive Capacity Armenia 4.74 Kazakhstan 1.15 Malaysia 0.54 Thailand 0.4 Azerbaijan 0.03

Source: Bauer et al. 2008.

Note: Based on $1.25 a day, which represents the international poverty line for extreme poverty. Poverty estimates are based on PPP for the year 2005.

No data could be found for most island countries. Anectodal data sources for indicate poverty levels above 30 percent for Afghanistan and Myanmar.

Page 260: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

260

Annex 5.1. Local coping strategies as adaptation tools to mitigate the impacts of climate change in agriculture.

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies

Non-Climate Benefits

Potential Mal-adaptation

Resources Required Source

Central

TAJIKISTAN Central Extreme cold

Shift of season

Loss of crops

Decreased food security

Improved cropping systems or alternative cultivation methods

Using cold frames to allow earlier seeding of plants

Higher income from farming

None Wood, glass and screws; seedbeds and watering

Total cost = US$ 90 per cold frame

Increased Growing Season in Central Tajikistan through Cold Frames. CARE Tajikistan. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=177)

Central (mountainous areas)

Shift of season

Extreme cold

Loss of crops

Loss of livelihoods

Improved food processing and storage

Increasing knowledge of food preservation and canning techniques to respond to winter food insecurity

Increased inter-community trade in fruits and vegetables

None Food preservation and canning equipment like glass jars or plastic bottles

Food Preservation and Canning in Mountain Communities of Central Tajikistan. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=178).

Erratic rainfall Landslides

Land degradation

Soil erosion

Afforestation/

Reforestation

Combining the planting of trees and an innovative watering system

Cultivation of leguminous as intermediate crops, mulching and use of compost as fertilizers

None Fencing materials (cement poles or armature, wire netting), seedlings, plastic bottles for drop irrigation; labor

Total = US$1,700 per ha

Reforestation/Afforestation to Prevent Soil Erosion and Land Slides in Tajikistan. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=179)

Central Floods

Extreme cold

Damage to forests

Land degradation

Soil erosion

Developing joint adaptation projects across communities

Dissemination of knowledge, education

Disaster risk management

Higher risk awareness None Maps, papers, shovels, poles

Total cost = US$1,200

Community Risk Assessment and Mapping in Central Tajikistan. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=181)

Page 261: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

261

Region/Country

Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies

Non-Climate Benefits

Potential Mal-adaptation Resources Required Source

Eastern

CHINA Western and Northern China of Yellow River (Loess Highlands)

Floods Soil erosion Check dams; controlling soil erosion

Controlling soil erosion through a series of dams or dam-fields

Higher crop yields Unable to treat the whole watershed for soil erosion particularly those occurring at the sides and top of the hills

Dam-fields suffer from salinization due to the high concentration of mineral salts in the water; upon evaporation, these salts were deposited on soil surface which are harmful to the crops; no specific measure designed to check salinization in the dam-fields

Locally available materials like stones, clay, pebbles

Tools and understanding of soil and water flows

Dam-fields in Northwest China. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=40)

Western Sichuan, Tibetan Plateau

Extreme cold Low survival/productivity of livestock

Appropriate livestock breeding

Livestock selection, e.g. breeding jiulong (valley-type) and maiwa (plateau-type) yaks

Enhances yaks and knowledge on breeding

None Yaks and knowledge on breeding

Ning Wu. 1998. Indigenous knowledge of yak breeding and cross-breeding among nomads in western Sichuan, China. In: IK Monitor Vol. 6(1)

Page 262: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

262

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

South Asia

BANGLADESH Jamalpur district Floods Loss of livelihoods

Livelihood diversification through integrated agriculture-aquaculture system

Establishing a community rice-fish farm

Increased community cooperation

Improved use of resources

Increased fish availability in the area

Increased income

None Training on organizational management

Fingerlings

Rice varieties for planting

Rice bran to feed fish

Bamboo, nets, tree branches, bushes

Shovels and other hand tools for digging

Making bunds and fish enclosures

Dey, M. and M. Prein. 2005. Increased Income From Seasonally Flooded Rice Fields through Community Based Fish Culture in Bangladesh and Viet Nam. Plant Prod Sci 8(3): 349-353

FAO 2001. Integrated agriculture-aquaculture. A primer. FAO Fisheries Technical Paper 407

K. M. Reshad Alam, M. C. Nandeesha, and Debasish Saha, Community Rice-Fish Farming in Bangladesh

Prein, M. and M.M. Dey. Rice and Fish Culture in Seasonally Flooded Ecosystems

Floods Loss of crops

Loss of livelihoods

Appropriate crop selection

Alternative cultivation methods or hydroponics

Adjusting transplanted Aman rice cultivation to more frequent floods

Higher crop yields None Follow the flood schedule

- establish early or late varieties of transplanted Aman (wet season rice) to avoid loss of crops due to variations in flood recurrence

- take advantage of the early production of rice by growing additional crops

Early or Late Transplanted Aman Rice Production in Bangladesh. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=194)

Livelihood Adaptation to Climate Variability and Change in Drought-Prone Areas in Bangladesh – DP9/1-BGD/01/004/01/99, Asian Preparedness Centre, Food and Agriculture Organization of the United Nations, Establishment of field demonstrations for Kharif II season/June-October 2007

Page 263: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

263

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-

adaptation Resources Required Source

BANGLADESH Floods Waterlogging Appropriate crop selection

Growing salient-resistant crops like Mele

Growing salient-resilient reeds to earn additional income

Potential for micro-enterprise (sold raw or as woven mats)

Species selection critical in Mele cultivation to ensure favorable results

Mele (Cyperus tagetiformus) seeds

Waterlogged area

Labor

Cultivation of Mele Reed in Bangladesh. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=82)

ITDG-B. 2003. An Attempt on Application of Alternative Strategies for Community Based Flood Preparedness in South-Asia, Bangladesh

Southwestern but applicable to other flooded areas of the country

Floods Waterlogging Alternative cultivation methods like hydroponics

Growing of crops or vegetables in floating gardens

Subsistence food during flooding

Potential source of additional income

None Seeds and seedlings

Water hyacinth

Paddy straw

Labor

Hydroponics in Bangladesh. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=80

Southwestern coastal area

Floods Low survival/productivity of poultry

Poultry breeding Raising duck during monsoon

Diet diversification

Food during monsoon

Cash generation for household needs

None Ducklings

Small shelter

Moderate technical knowledge

Access to waterbodies, vaccines

Small quantity of supplementary feed (locally available fish)

F. Mallik 2005 Adaptation in Action. Community Level Adaptation Techniques in the Context of the Southwestern Region of Bangladesh

ITDG-B 2003 An Attempt on Application of Alternative Strategies for Community Based Flood Preparedness in South-Asia, Bangladesh

Page 264: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

264

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies

Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

BANGLADESH Sea-level rise Loss of crops Appropriate crop selection

Cultivating maize and fodder grass during dry season

Increased paddy yield

Livestock raising as alternative livelihood

None Seeds and seedlings

Livestock

Labor

Hossen and Roy 2005. Local Contributions to Operationalising the UNFCCC, CBD and UNCCD. Reducing Vulnerability to Climate Change in the Southwest Coastal Region of Bangladesh

Northwestern district

Barind tract called as Varenda Tract; includes Dinajpur, Rangpur, Pabna, Rajshahi, Bogra and Joypurhat districts of Rajshahi Division

Drought/aridity Loss of crops

Water shortage

Appropriate crop selection

Alternative cultivation methods

Planting drought-resistant fruit trees to secure income:

Mango gardening in NW

Jujube gardening in Barind tract

Households with alternative source of livelihood

Increased farmer income

Ensures food security

Diet diversification

3-4 year old mango trees produce a high shadow cover that threatens rice if used for intercropping, since the latter does not grow under the shadow

Under changing climatic conditions, increasing temperatures may induce synchronized maturity which could lead to price drops

Seedlings of fruit trees

R. Selvaraju, A.R. Subbiah, S. Baas, I, Juergens, Livelihood Adaptation to Climate Variability and Change in Drought-Prone Areas in Bangladesh – Case Study, implemented under the Project Improves Adaptive Capacity to Climate Change for Sustainable Livelihoods in the Agriculture Sector – DP9/1-BGD/01/004/01/99, Asian Preparedness Centre, Food and Agriculture Organization of the United Nations, Rome, 2006

Northwestern

Barind tract

Drought/aridity Water shortage Improved cropping system through alternative cultivation method

Adjusting transplated aman seeding practices to more frequent droughts

Adopt alternative seedbed methods for timely transplanting of seedlings; these methods may be mat-type seedlings in tray, dry seedbeds; dapog nurseries

Higher crop yields None Materials for seedbed

Rice seedlings

Jensen, J.R., Mannan, S.M.A., Uddin, S.M.N. 1993. Irrigation requirement of transplanted monsoon rice in Bangladesh, Agricultural Water Management, 23: 199-212.

Page 265: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

265

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits

Potential Mal-adaptation

Resources Required Source

BANGLADESH Drought/aridity Water shortage Rainwater harvesting Using gutters and pipes to collect rooftop water

Improved health and sanitation

Boost local enterprise

None Gutters

Pipes

Storage tank

UNEP DTIE (2000), Rooftop Rainwater Harvesting for Domestic Water Supply. In: Sourcebook of Alternative Technologies for Freshwater Augmentation in Some Countries in Asia, UNEP DTIE. (http://www.unep.or.jp/ietc/publications/techpublications/techpub-8e/rooftop.asp)

Drought/aridity Land degradation Alternative cultivation methods

Using organic matter to enhance soil

Higher crop yield None Homestead waste

Water hyacinth

Debris

Dung

Farm yard

Two Chamber Farm Yard Manure/Water Hyacinth Compost Preparation in Bangladesh. UNFCCC case study. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=191)

Amoding, A., Muzira, N.R., Bekunda, M.A and Woomer, P.L. 1999. Bioproductivity and decomposition of water hyacinth in Uganda. African Crop Science Journal 7:433-439

Drought/aridity Land degradation

Soil erosion

Soil conservation

Livelihood diversification

Home gardening as a means to climate- proofing farming

Increased farmer’s income

None Vegetable seeds

Seeds/seedlings of drought tolerant tree and vegetable species

Backyard as garden

Homestead Gardens in Bangladesh. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=192)

Homestead food production- an effective integrated approach to improve food security among the vulnerable char dwellers in northern Bangladesh. Homestead Food Production Bulletin No. 4. Dhaka: Helen Keller International/Bangladesh

Page 266: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

266

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

BHUTAN Wangling, Jangbi, Phumzur villages in Trongsa district

Erratic rainfall Loss of crops Diet diversification Harvesting wild vegetables, fruits and tubers from the forest by the Monpas, a Bhutanese ethnic group

Ensured food security

Indigenous knowledge passed from generation to generation preserves this practice

None Knowledge about flora and fauna

Labor and skills for collecting wild edibles

Harvesting Wild Foods in Bhutan. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=7)

Centre for Bhutan Studies (http://www.bhutanstudies.org.bt/main/index.php0

West central Drought/aridity Water shortage Rainwater harvesting

Collecting, storing and distributing water through a tank system

Vegetable production None Materials to construct water tanks, bamboo or polythene pipe for water distribution

Some maintenance is required periodically to ensure collection and equal distribution of water to the beneficiaries

Water Storage Tanks in Bhutan. UNFCCC database on local coping strategies.. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=26)

Drought/aridity Loss of crops

Water shortage

Sustainable water management

Using bamboo stems for drip irrigation during the dry season

Increase in crop yield None Bamboo and local labour for setting up the drip-irrigation system and filling the bamboo with water regularly during the dry period; bamboo needs to be replaced after five years or more

Drip Irrigation in Bhutan. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=21)

Down to Earth 2003 (http://www.cseindia.org/dte-supplement/water-index.htm)

Drought/aridity Loss of crops Alternative cultivation methods

Managing common pool resources

Preservation of traditional agriculture system practiced by small farmers

None Labor

Seeds

Livestock

Access to forest resources

Integrated Farming systems in Bhutan. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=161)

Sonam Tobgay. 2005. Small Farmers and the Food System in Bhutan, Agricultural Marketing Services, Ministry of Agriculture, Royal Government of Bhutan.

Page 267: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

267

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

HIMALAYAS Drought/aridity Water shortage

Soil erosion

Rainwater harvesting

Soil conservation

Controlling soil erosion and managing rainwater

- Terracing

- Field leveling

- Ploughing

- Sheet erosion control

- Wind erosion control

- Biofencing

Higher crop yields Stones, gravel and boulders Grasses, bamboo, shrubs

and seabuckthorn (Hippophae rhamnoides)

Shovels, tamping tools Plough Labor

L.R. Verma 1998. Soil and water management techniques In: Indigenous technology knowledge for watershed management in upper north-west Himalayas of India, PWMTA Field Document No. 15, Kathmandu.(http://www.fao.org/docrep/X5672E/x5672e03.htm

INDIA Uttar Pradesh Floods Loss of crops Appropriate crop selection

Breeding rice varieties in flood-prone areas

Higher crop yields

Ensured food security

Agricultural expertise on flood-prone rice and collection methods

Rice seeds Partnerships between

scientists and farmers

D.M. Maurya . 1997. Participatory Breeding, On-farm Seed Management and Genetic Resource Conservation Methodology (http://archive.idrc.ca/library/document/104582/maurya.html)

J.L. Dwivedi 1997. Conserving genetic resources and using diversity in flood-prone ecosystems in eastern India (http://www.idrc.ca/en/ev-85297-201-1-DO_TOPIC.html)

Himachal Pradesh

Erratic rainfall Water shortage Sustainable water management

Utilizing and distributing glacier runoff

Improved agricultural output and food security

Improved health

If a similar system is indented to be set up, one has to pay due attention to water rights

Rocks Wood Tools Cement Pipes

L.R. Verma 1998. Soil and water management techniques In: Indigenous technology knowledge for watershed management in upper north-west Himalayas of India, PWMTA Field Document No. 15, Kathmandu.(http://www.fao.org/docrep/X5672E/x5672e03.htm)

Waterharvesting.org Kul Irrigation of the Trans-Himalaya (http://www.rainwaterharvesting.org/methods/traditional/kuls.htm)

Arunachal Pradesh

Erratic rainfall Loss of crops Appropriate crop selection

Domesticating indigenous varieties of cereals and fruit trees

Promotes local enterprises for women

None Seeds and seedlings Labor

Diversifying Crops in Arunachal Pradesh, India. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=79)

Page 268: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

268

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

INDIA Himalayas Erratic rainfall Loss of crops Alternative cultivation methods

Growing apricots, walnuts, grapes and vegetables in the cold deserts

Farmer’s practices to cultivate fruits and vegetables to ensure stable supply of vitamins:

- Root spreading of cabbage through use of tokhre (small wooden structure) for surface feeding of nutrient or moisture uptake

- Mahotra/dhing/guchhi (mushroom) harvested in grasslands or forests for food and additional income

- Localized greenhouse grape cultivation in Nubra valley

- Apricot grafting

- Fruiting walnuts

Higher crop yield None Indigenous knowledge for recognizing mushroom species

Bricks

Knives or other cutting tools

Walnut and apricot trees

L.R. Verma. 1998. Indigenous technology knowledge for watershed management in upper north-west Himalayas of India PWMTA Field Document No. 15, Kathmandu: FAO (http://www.fao.org/docrep/X5672E/x5672e00.htm)

Erratic rainfall Loss of crops Appropriate crop selection in cold deserts

Rotational cropping

Seed selection

Improved soil properties

Higher crop yields

None Seeds for appropriate species for intercropping

Wooden tools for spreading soil

L.R. Verma. 1998. Indigenous technology knowledge for watershed management in upper north-west Himalayas of India PWMTA Field Document No. 15, Kathmandu: FAO (http://www.fao.org/docrep/X5672E/x5672e00.htm)

Page 269: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

269

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

INDIA Western Himalayas Erratic rainfall Loss of crops Disaster risk management

Appropriate cropping practices

Using meteorological indicators and animal behavior to predict rain such as

- Visible spectrum around the sun and moon

- Clouds and wind direction

- Activities of various birds

- Animals and insects

- Crop performance

- Condensation

Possibility for reduced livelihood losses

As climate change occurs, these traditional forecasting indicators may change. Locals have to continue their observations and adjust their predictions accordingly to ensure that correct coping mechanisms will be applied

Indigenous forecasting knowledge

Indigenous Forecasting in Western Himalayas. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=46)

L.R. Verma. 1998. Indigenous technology knowledge for watershed management in upper north-west Himalayas of India PWMTA Field Document No. 15, Kathmandu: FAO (http://www.fao.org/docrep/X5672E/x5672e00.htm

Himalayas Erratic rainfall Water shortage Rainwater harvesting Using roofs, ponds and tanks to harvest rain, dew and fog water

Higher crop yields Tools

Materials to build an irrigation systems

Cement or pang grass to line storage areas

L.R. Verma. 1998. Indigenous technology knowledge for watershed management in upper north-west Himalayas of India PWMTA Field Document No. 15, Kathmandu: FAO (http://www.fao.org/docrep/X5672E/x5672e00.htm

Himalayas Erratic rainfall Loss of crops Appropriate crop selection Rotational cropping

Seed selection

Improved soil properties

Higher crop yields

Bibe Seeds of appropriate species for intercropping

Wooden tools for spreading soil

L.R. Verma. 1998. Indigenous technology knowledge for watershed management in upper north-west Himalayas of India PWMTA Field Document No. 15, Kathmandu: FAO (http://www.fao.org/docrep/X5672E/x5672e00.htm

Page 270: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

270

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

INDIA Goa Sea-level rise Waterlogging Integrated agriculture-aquaculture system

Balancing agriculture and fisheries through sluice gates

Application of khazan – traditionally community managed integrated agriculture-aquaculture ecosystems

Promotes symbiotic relationship between the rich and poor class through employment generation and labor sharing

None Agricultural land

Labor

Wood for shutters of sluice gates

Canoes and nets for fishing

TERII, INTEREST - Interactions between Environment, Society and Technology. Three case studies using different ecosystems: Traditional Aquaculture – Goa, Agriculture – Karnataka, Bamboo forests – Haryana. The Energy and Resources Institute (http://www.teriin.org/teri-wr/projects/interestaqua.pdf)

Drought/aridity Water shortage Rainwater harvesting Building anicuts (small-medium sized dams) to serve as water reservoirs such as

- Supplementary irrigation during erratic monsoons

- Groundwater recharge during very low water levels

Increased crop yield

Sustainable supply of drinking water for cattle and people

Better hygiene – bathing of men and women

None Stone

Mud

Concrete or local materials to construct dam

labor

Narain, P.; Khan, M. A.; Singh, G. 2005, Potential for water conservation and harvesting against drought in Rajasthan, India. Working Paper 104 (Drought Series: Paper 7). Colombo, Sri Lanka: International Water Management Institute (IWMI).

Anicuts in India. UNFCCC case study (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=12)

Himalayas Erratic rain

Drought/aridity

Land degradation Nutrient management Manure and ash application to increase soil fertility

- Organic manure

- Crop residue harvesting

- Kitchen ash

Improved soil properties

Higher crop yields

None Farm yard manure

Compost

Tools

L.R. Verma. 1998. Indigenous technology knowledge for watershed management in upper north-west Himalayas of India PWMTA Field Document No. 15, Kathmandu: FAO (http://www.fao.org/docrep/X5672E/x5672e00.htm

Page 271: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

271

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-

adaptation Resources Required Source

INDIA Central Himalayas (northwest most border areas), Garhwal region

Drought/aridity Loss of crops Diet diversification Use of wild foods and medicinal plants by Bhotiya tribes (Tolchha, Marchha, Jadhs)

Improved health and nutrition

None Local knowledge about plant resources

R.K. Maikhuri, Sunil Nautiyal, K.S. Rao and R.L. Semwal 2000. Indigenous knowledge of medicinal plants and wild edibles among three tribal subcommunities of the Central Himalayas, India. In: IK Monitor Vol. 8(2) (http://web.archive.org/web/20041215132544/www.nuffic.nl/ciran/ikdm/8-2/maikhuri.html)

Thar Desert Drought/aridity Water shortage Rainwater harvesting Building underground tanks (called as kunds) for collecting and storing water

Improvements in health, i.e., reduced water-borne diseases which are common in desert areas

None Lime plaster

Building materials

Gravel, pond silt or charcoal ash

Wire mesh

Tools

Rainwater harvesting. (http://www.rainwaterharvesting.org/methods/traditional/kunds.htm)

Kunds in Thar Desert, India. UNFCCC database on local coping strategies.. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=57)

Northeast Drought/aridity Loss of crops

Water shortage

Sustainable water management

Using bamboo to transport stream and springwater to irrigate plantations by the Meghalaya tribal farmers

Higher crop yields None Bamboo for pipes and stakes

Tools

Local network for maintenance of the bamboo irrigation system

Drip Irrigation in Northeast India. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=58)

Andaman and Nicobar Islands

Drought/aridity Loss of crops Alternative cultivation method

Intercropping with banana and using plant residues

Higher crop yields None Seedlings

Plant residues

Dr. A.K. Bandyopadhyay, Director (CARI) and Dr. G.S. Saha, Scientist (CIFA), Coping with heat and water shortages on the Andaman and Nicobar Islands, India, Indigenous Knowledge and Development Monitor, Vol 7(2), July 1999, pp. 26-27

Page 272: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

272

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-

adaptation Resources Required Source

INDIA Andaman and Nicobar Island

Drought/aridity Loss of crops Appropriate crop selection

Selecting and storing rice, pulse and vegetable seeds

Selected seeds provide higher yields

None Seeds and containers to store dried seeds

Leaves of Neem or Salamu plant to protect from insects

Cow dung

A.K. Bandyopadhyay and G.S. Saha (1998): Indigenous methods of seed selection and preservation on the Andaman Islands in India. In: IK Monitor Vol. 6(1) (http://web.archive.org/web/20041215132629/www.nuffic.nl/ciran/ikdm/6-1/bandy.html)

Gujarat Drought/aridity Water shortage Rainwater harvesting De-sitting, cleaning and deepening of ponds to collect rainwater

Increased crop yield

Reduced labor migration

None Plastic lining for ponds

Labor to dig and deepen wells and ponds

Ponds in Gujarat, India. UNFCCC database on local coping strategies (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=16)

Maharashtra Drought/aridity Water shortage

Soil erosion

Rainwater harvesting Building ground barriers and shallow excavations through

- Various barriers like contour bunds, nalla bunds, check dams, gabions

- Shallow excavations like contour trenches, farm ponds, reservoirs in bedrock

- Roof tops

- Water recycling by using domestic waste water to irrigate kitchen gardens

Improved food security and livelihoods

Increased understanding and educational opportunities for the community on water resource management

None Training on water harvesting and resource management

Expertise on hydrology and hydrogeology

Tools

UNESCO, Conjunctive use of water resources in Deccan Trap. In UNESCO Best Practices on IK. (http://www.unesco.org/most/bpik13-2.htm)

R.K. Sivanappan 1997. Technologies for water harvesting and soil moisture conservation in small watersheds for small-scale irrigation, FAO

Page 273: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

273

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

INDIA Orissa Drought/aridity Loss of crops Appropriate crop selection

Storing and exchanging rice varieties and medicinal plants

Higher crop yields

Ensured food security

Increase in knowledge of indigenous varieties

Higher biodiversity

Communities and projects must ensure that the vegetables, trees and other plants promoted are suitable in the areas

Seedlings

Knowledge on how/when to plant and preserve them

Farmers network

Storage facilities

Seed Banks in Gujarat, India. UNFCCC database on local coping strategies (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=59)

Drought/aridity Water shortage Rainwater harvesting Water management practices including conservation, recycling, instilling and maintaining tube wells

Better health and hygiene

Promotes cohesiveness within the communities

Acts as social safety nets in times of disaster

None Human resources

Materials for construction of tube wells

Sanjoy Bandyopadhyay 2003. Coping strategy and vulnerability reduction to Climate Extremes, Presentation at the Expert workshop on local coping strategies and technologies for adaptation, Delhi, India. (http://unfccc.int/files/meetings/workshops/other_meetings/application/pdf/sanjoy.pdf)

Rajasthan Drought/aridity Water shortage Rainwater harvesting Harvesting water and recharging groundwater through earthen check dams (johads)

Higher crop yields

Increase forest coverage

Associated availability of fuelwood and tree leaves for fodder

None Concrete

Soil

Shovels

Buckets

Labor

Johads in Rajasthran. UNFCCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=41)

Drought/ariditiy Water shortage Rainwater harvesting Building contour bunds (contour ridges) to collect water run-off

Reclamation of degraded land

None Stone

Mud

Concrete or other local materials to construct bunds

Labor

Bunds in Rajasthan, India. UNFCCC database on local coping strategies (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=9)

Drought/aridity Loss of crops Appropriate crop selection

Cultivating Bajra millet in arid regions (millet can be cultivated in sandy and under rainfed conditions)

Food security None Seeds and seedlings

Labor

Bajra Millet in Rajasthan, India. UNFCCC database on local coping strategies (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=13)

Page 274: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

274

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits

Potential Mal-adaptation

Resources Required Source

INDIA Rajasthan Drought/aridity Loss of crops Appropriate crop selection

Income diversification

Growing “Sona Muhi” (Cassia angustifolia) as medicinal cash crop

Cash income None Sona-mukhi seedlings

Labor

Access to markets

Brook Bhagat 2002. Spitting in the wind - Combating Desertification in the Great Indian Desert (http://ecoworld.com/Home/Articles2.cfm?TID=323)

Drought/aridity Land degradation Nutrient management

Using worms to process organic waste

Increase in crop yield

Reclamation of degraded land

None Organic waste

Earthworm

labor

Dr. Henamgee Jambhekar, Vermiculture in India (http://fadr.msu.ru/rodale/agsieve/txt/vol7/art3.html)

Tamil Nadu Drought/aridity Water shortage Sustainable water management

Improving wells and irrigation Improved agricultural outputs

Improved livelihoods

None Tools

Pipes

Lime mortar

Technical expertise for siting, construction and maintenance measures

M. Manoharan and S. Kombairaju 1995. ITK suits transported sandy soils. In: IK Monitor Vol. 3(1). (http://web.archive.org/web/20041217210736/www.nuffic.nl/ciran/ikdm/3-1/articles/manoharan.html)

Drought/aridity Soil erosion Soil conservation Applying soil amendments to improve soil properties

Improved agricultural outputs

Improved livelihoods

Possible salinization

Tank silt

Machinery and tools

Pipes

Lime mortar,

Technical expertise for siting, construction and maintenance measures

M. Manoharan and S. Kombairaju (1995): ITK suits transported sandy soils. In: IK Monitor Vol. 3(1) (http://web.archive.org/web/20041217210736/www.nuffic.nl/ciran/ikdm/3-1/articles/manoharan.html)

John Butterworth, Barbara Adolph and Suresh Reddy (2003) How farmers manage soil fertility: a guide to support innovation and livelihoods. Chapter 4 Soil amendments Hyderabad: Andhra Pradesh Rural Livelihoods Project/Chatham: Natural Resources Institute

Drought/aridity Water shortage

Soil erosion

Soil conservation Coping with wind erosion through - Application of farm yard

manure - Use of coconut - Planting drumstick

species (Jafna or Yalpanam Murungai)

Higher yields

Additional income

None Farm yard manure

Coconut seedlings less than six months old

Banana plants

10% dust of benzene hexachloride

Diammonian phosphate complex

Potash mixture

Stem cuttings of 21/2-3 ft of drumstick

Pruning shears or knives

M. Manoharan and S. Kombairaju 1995. ITK suits transported sandy soils. In: IK Monitor Vol. 3(1) (http://web.archive.org/web/20041217210736/www.nuffic.nl/ciran/ikdm/3-1/articles/manoharan.html)

Page 275: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

275

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-

adaptation Resources Required Source

INDIA Tamil Nadu Drought/aridity Loss of crops Post-harvest management

Threshing, winnowing, cleaning and drying for dryland crops

Improved nutrition

Recognition of women farmers for their knowledge and practice

Important to consider local humidity and pests for replication of this post-harvest technology

Labor

Tools and skills required are simple

S. Parvathi, K. Chandrakandan and C. Karthikeyan 2000. Women and dryland post-harvesting practices in Tamil Nadu, India. In: IK Monitor Vol. 8(1) (http://web.archive.org/web/20041204232430/www.nuffic.nl/ciran/ikdm/8-1/parvathi.html)

Uttar Pradesh Drought/aridity Land degradation Nutrient management Increasing soil fertility through gypsum, manure and compost applications

Reclamation of degraded soil

Increase in crop yield

None Labor

Seeds and seedlings

Livestock for manure

Organic pesticides such as Neem

Department of Agriculture and Cooperation (2005), Uttar Pradesh Sodic Land Reclamation Project with World Bank Assistance (Phase II), Indian Ministry of Agriculture (http://agricoop.nic.in/PolicyIncentives/nrmd.htm)

NEPAL Extreme cold Loss of crops Post-harvest management

Processing green leafy vegetables

Promotes local enterprise for mountain women

None Green leafy vegetables

Wooden stick to beat the vegetables

Containers to store vegetables

Narayan P. Manandhar 1998. The preparation of gundruk in Nepal. A sustainable rural industry? In: IK Monitor Vol. 6(3) (http://web.archive.org/web/20041215133649/www.nuffic.nl/ciran/ikdm/6-3/manandh.html)

PAKISTAN Sindh Drought/aridity Water shortage Sustainable water management

Building laths at different levels to irrigate fields

Increase in crop yield Possible land degradation as some areas are no longer flooded

Community water management system

Machinery to build canals and move earth

Spate Irrigation in Sidh, Pakistan. UNFCCC database on local coping strategies. (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=69)

Page 276: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

276

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-adaptation

Resources Required Source

SRI LANKA Drought/aridity Water shortage Rainwater harvesting (cascading tanks)

Using stored water efficiently

Increased crop yield Overexploitation of tanks and poor management disturb water use within watersheds

Bricks and cement

Tools

Labor

Tikiri Nimal Herath (2001): Farmer's knowledge of water management methods in the dry zone of Sri Lanka. In: IK Monitor Vol. 9(3) (http://web.archive.org/web/20040719195342/www.nuffic.nl/pdf/ciran/ikdm0111.pdf)

Drought/aridity Water shortage Rainwater harvesting Managing water by women

Improved health through clean water supplies

Improved food security

Diversified diets

None Tanks

Pots

Styrclinos potatorum seeds

Training or public awareness on water conservation/efficiency

R.K. Ulluwishewa (1994): Women's indigenous knowledge of water management in Sri Lanka. In: IK Monitor Vol. 2(3) (http://web.archive.org/web/20041218040330/www.nuffic.nl/ciran/ikdm/2-3/articles/ulluwishewa.html)

Anuradhapura District

Drought/aridity Water shortage Sustainable water management

Distributing the work of maintaining and repairing small-scale irrigation systems (the practice is called as pangu)

Improved crop yield None Tools and labor

Community organization

Pangu Practice in Sri Lanka. (http://www.unesco.org/most/bpik22.htm)

Drought/aridity Loss of crops Alternative cultivation methods

Zero-tillage paddy cultivation (Nawa Kekulama)

Increase crop yield None Labor for the paddy cultivation

G.K.Upawansa 1997 New Kekulam rice cultivation: a practical and scientific ecological approach. In LEISA Magazine, volume 13, Issue 3 - Rebuilding Lost Soil Fertility

Drought/aridity Loss of crops Land redistribution Temporary redistribution of private fields called as bethma practice. It temporarily redistributes plots of land among shareholders (being paddy landowners) in part of the command area of a tank (water reservoir) during drought periods

Higher crop yields None Some leadership to prevent/solve conflicts and ensure the bethma practice functions well

UNESCO, The bethma practice: promoting the temporary redistribution of lands during drought periods. UNESCO Best practices on IK (http://www.unesco.org/most/bpik21.htm)

Region/Country Local Area Natural Disaster Impacts Adaptation Action Local Coping Strategies Non-Climate Benefits Potential Mal-

adaptation Resources Required Source

Page 277: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

277

SRI LANKA Drought/aridity Loss of crops Pest control Controlling weed growth through dry straw in paddy fields

Increased crop yields None Paddy seeds

Banana plant

labor

Control of Weed Growth in Sri Lanka. UNFCCC database on local coping strategies (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=30)

SOUTH EAST ASIA

INDONESIA West Java Drought/aridity

Floods

Loss of crops Alternative cultivation methods

Growing fish on huma (dry swidden fields) and sawah (wet fields)

Generation of cash income

None Labor

Paddy fields

Paddy seeds or seedlings

Fish or fingerlings

FAO (2001), Integrated agriculture-aquaculture. A primer. FAO Fisheries Technical Paper 407 (http://www.fao.org/docrep/005/y1187e/y1187e00.htm)

TIMOR Erratic rainfall

Storms

Loss of crops Appropriate crop selection

Strategies for seed selecting and plating to cope with the disasters

Reduced agricultural losses

Improved food security

Increased biodiversity

None Seeds variety

Understanding of breeding and weather forecasting

Johan Kieft 2001. Indigenous variety development in food crops strategies on Timor. In: IK Monitor Vol. 9(2) (http://web.archive.org/web/20041221071223/www.nuffic.nl/ciran/ikdm/9-2/kieft.html)

GREATER MEKONG SUBREGION

LAO PDR Attapeu province Floods Loss of crops Alternative cultivation methods

Diet diversification

Diversifying rice-based diets during flood season

Improved health from diversified diets

Prolonged food shortages threaten wetland and forest resources

Fishing equipment and expertise

Meusch, E., Yhoung-Aree, J., Friend, R. & Funge-Smith, S.J. (2003) The role and nutritional value of aquatic resources in the livelihoods of rural people – a participatory assessment in Attapeu Province, Lao PDR. FAO Regional Office Asia and the Pacific, Bangkok, Thailand, Publication No. 2003/11

MEKONG DELTA Floods

Sea level rise

Storms

Loss of crops

Loss of land

Damage to human settlements

Disaster risk management

Building forecasting capacity and adaptation strategy

Building Forecasting and Capacity and Adaptation Strategy in the Mekong Delta. UNFCCC database on local coping strategies (http://maindb.unfccc.int/public/adaptation/adaptation_casestudy.pl?id_project=197)

Page 278: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

278

APPENDIX

Appendix 1. Country case studies

Climate Change and Agricultural Development: Case Study in Bangladesh

Introduction

Bangladesh is one of the largest deltas in the world, made up of a dense network of the tributaries of the rivers Ganges, the Brahmaputra and Meghna (Bass and Ramasamy 2007). The agricultural sector accounts for 20 percent of the GDP and employs 52 percent of total labor force (World Bank 2008). Bangladesh has experienced economic growth in the last decade and has moved from the low human development index category to a medium human development index category. However, around 36 percent of the population still lives on a dollar per day or less, 80 percent on less than two dollars a day and almost 40 percent of children under five are underweight (Government of Bangladesh 2005; OneWorld 2008; Ansorg and Donelly 2008). Bangladesh grows only about 90 percent of its rice it needs in a typical year, which makes it the world’s fourth largest rice importer (OneWorld 2008).

Natural disasters are frequent in the country, mainly floods, cyclones and storm surges, flash floods, droughts, tornados, earthquakes, riverbank erosion, and landslides. The low lying coastline of Bangladesh makes it highly vulnerable to sea level rise. About 80 percent of the country’s land is affected by floods which in a normal year inundates 20 to 25 percent of the country through river spills and drainage congestions (Government of Bangladesh 2005). Furthermore, about 1.7 million hectares of floodplain areas are prone to riverbank erosion and around 42 percent of the coastal and offshore arable land areas are affected by different degrees of soil salinity. Main causes of salinization are tidal flooding during the wet season, direct inundation by saline or brackish water and upward or lateral movement of saline groundwater during the dry season, and inundation with brackish water for shrimp farming (Government of Bangladesh 2005).

The depletion of land and water resources associated with a growing food demand increases the threats of climate change to food production and security in Bangladesh. During the boro season (rice monocropping), land degradation is caused by chemical fertilizers, pesticides and lack of crop diversification. Overexploitation of groundwater for irrigation is also causing reduction in groundwater aquifers (Selvaraju et al. 2006).

Impacts on Agriculture

In Bangladesh, crop production is constrained by too much water during the wet season and too little during the dry season (Government of Bangladesh 2005). Any changes in mean temperature, rainfall and in the intensity of climate extremes are likely to aggravate current vulnerabilities of the agricultural sector. Scenarios project that by 2050, dry season rainfall may decrease by 37 percent and monsoon rainfall might increase 28 percent. Therefore, an increase of droughts in the dry season and floods in the rainy season is expected, which, added to encroaching salinity due to rising sea levels, will further degrade agricultural areas (Bass and Ramasamy 2007).

Wheat production is likely to be severely affected by climate change under different concentrations of CO2 fertilization, particularly, above temperature increases of 4oC. Impacts on

Page 279: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

279

rice production will be highly dependent on the extent of temperature increase and if the beneficial effects of CO2 fertilization take place (Karim, Hussain, and Ahmed 1999). However, the inclusion of multiple stressors such as moisture stress, inundation, and salinity in climate projections leads to substantial decreases in rice and wheat production under temperature increases of 4oC, even considering the beneficial CO2 effects (Faisal and Parveen 2004). More studies should be conducted to assess the impacts of climate change and rising sea levels on agriculture and rural communities in Bangladesh, especially the impacts of more frequent and intense floods and droughts.

Coping and Adaptation Strategies

In Islampur Upazilla in the Brahmaputra River basin, communities often develop strategies to cope with floods. However, during a survey, the great majority of respondents reported that they adopt short-term (relocation to safer places) or at most medium-term (store grains in safer places) measures to cope with the hazard. Only a handful of households considered longer-term measures such as building dwellings on stilts ( Mirza 2007). As a result of the recurrent droughts, farmers also have to adapt their farming systems from year to year according to the conditions of droughts. However, most do not have the resources for agricultural adjustments such as re-sowing, crop replacement, intercropping or irrigation. Therefore, most farmers rely on disposal/mortgaging of assets, borrowing and migration. In a case study in 1994-95, 72 percent of households of a community sold and/or mortgaged their lands as drought coping mechanisms, which led farmers into a debt trap (Bass and Ramasamy 2007). Such measures are also insufficient for proper adjustment to future climate variability and change (Selvaraju et al. 2006).

A study in Chapai Nawabganj (Gomastapur and Nachole Upazillas) and Naogoan (Porsha and Sapahar Upazillas) shows the following coping strategies in agriculture during drought events (Selvaraju et al. 2006):

i) traditional locally-managed practices such as pond excavation, retention of rainwater in khari canals and moisture conservation

ii) government-supported practices such as deep tubewell facilitated irrigation, supplemental irrigation and miniponds

iii) alternative and innovative automatic adaptation practices such as adoption of mango farming, integrated crop-livestock farming systems

iv) technology driven efforts such as new short-duration and drought-tolerant crop varieties, cropping systems and homestead gardening

v) disposal of productive assets and mortgaging lands, and vi) institutional-support activities including support from the government, NGOs and

community-based organizations (CBOs).

In the Sunderbans of Bangladesh, millions of people are well adapted to tidal and seasonal variations in water and salinity levels. Local farmers cultivate flood-tolerant rice during the monsoon and harvest salt-tolerant fish during the dry season. However, climate change will increase the pressure on those traditional livelihoods, adding to the impacts of rapid population growth, poaching of wildlife, increased deforestation as a result of growing industrial demand and shrimp farming. Adaptation strategies will have to consider both climatic and non-climatic pressures ( Kelkar and Bhadwal 2007).

Page 280: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

280

A combination of natural and socio-economic vulnerabilities, geographic factors and weak infrastructure has been a source of insecurity for people in the country, reducing their ability to cope with recurrent floods and droughts (Ansorg and Donelly 2008). However, some good adaptation strategies can be found in the country. For instance, Oxfam and its local partner organization, Samaj Kallyan Sangsthan introduced new varieties of bean and papaya in the Khonchapara village in the Gaibandha district, which can be harvested after floods when paddy crops are destroyed. The new crops helped farmers to survive during the crisis period ( Kelkar and Bhadwal 2007).

In Subarnabah village in south-west Bangladesh, development organizations21 joined efforts to implement a project called Reducing Vulnerability to Climate Change (RVCC). The project’s purpose is to promote livelihood strategies for income and food generation which include goat, duck, and hen rearing, chicken and crab farming, tree planting, introduction of salt-water tolerant vegetable gardens and handicraft production in a number of flood-prone villages in costal Bangladesh. Villagers also have access to loans to establish small crab farming enterprises. The initiative is still recent but farmers have already begun to make some profit and are encouraging others to join the project ( Kelkar and Bhadwal 2007). It has also generated knowledge about climate change impacts and the need for adaptation measures at the community-level (Government of Bangladesh 2005).

Considering the frequent climate extreme events in Bangladesh, one of the key steps to improve the adaptive capacity of local communities is to improve access to information. A survey in a rural community showed that local people are not aware of flood warnings issued by the Flood Forecasting and Warning Centre (FFWC) located in Dhaka ( Mirza 2007). Therefore, flood warning systems and their dissemination in an accessible language to communities are essential.

Furthermore, adaptation policies should be context-specific. For instance, project demonstrations show that mini-ponds are good for a farmer operating on a clay soil but it is not a good adaptation practice for a farmer operating on sandy soil (Bass and Ramasamy 2007). Adaptation practices need also to be monitored; and risk of mal-adaptation should be alerted (Bass and Ramasamy 2007). In North West Bangladesh, landowners started to plant mango trees in their rice fields because of economic gain and its adaptability to dry conditions. This autonomous adaptation is likely to have negative impacts as a result of shading on the rice crop underneath in 2-3 years time (Bass and Ramasamy 2007). Some adaptive measures might also lead to conflict. For instance, farming shrimp can be a better way to cope with climate change than growing crops. However, most farmers do not have resources to make the initial investments and as a result they end up selling their salinized land to commercial enterprises that grow shrimp for export. It is estimated that for every 20 people employed in a 20 “bigha” paddy field, only one is employed in a shrimp farm of the same size. In Khulna and Satkhira Districts, there are cases of violent conflicts between local farmers who regret selling their lands and shrimp workers (Ansorg and Donelly 2008). Tensions arriving from temporary or permanent migration as a result of increasing flooding, rising water salinity and loss of land are also expected to increase.

Uncertainty about the impact of climate change cannot justify inaction (Bass and Ramasamy 2007). Due to Bangladesh’ high poverty rates and large agricultural sector, projected

21 The project is implemented by the Institute of Development Education for the Advancement of the Landless

(IDEAL) and CARE Canada through CARE Bangladesh, funded by the Canadian International Development Agency (CIDA)

Page 281: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

281

increases in climate extremes and sea level rise will significantly impact its vulnerable population. As mentioned above the identification of additional good practices, broader replication and exchange of current good practices are ways to decrease the vulnerability of the population.

References

Ansorg, T., and T. Donelly. 2008. Climate Change in Bangladesh: Coping and conflict. European Security Review no. 40Brussels: ISIS Europe

Bass, S., and S. Ramasamy. 2007. Improved Adaptive Capacity to Climate Change for Sustainable Livelihoods in the Agriculture Sector Summary Report. Project Phase I. Community Based Adaptation in Action . Rome: FAO

Faisal, I. M., and S. Parveen. 2004. Food security in the face of climate change, population growth, and resource constraints: Implications for Bangladesh. Environmental Management 34 (4): 487–98.

Government of Bangladesh. 2005. National Adaptation Programme of Action (NAPA). Dhaka: Ministry of Environment and Forests

Karim, Z., G. Hussain, and A. U. Ahmed. 1999. Climate change vulnerability of crop agriculture. In Decision criteria and optimal inventory processes, ed. B. Liu and A. O. Esogbue. New York, Springer.

Kelkar, U., and S. Bhadwal. 2007. South Asian Regional Study on Climate Change Impacts and Adaptation: Implications for Human Development. Occasional Paper No. 2007/27. UNDP

Mirza, M. 2007. Climate Change, Adaptation and Adaptive Governance in Water Sector in South Asia. Toronto: Environment Canada

OneWorld. 2008. Hunger Countries at Risk. OneWorld.net http://uk.oneworld.net/guides/food/countries (accessed February 3, 2009)

Selvaraju, R., A. R. Subbiah, S. Bass, and I. Juergens. 2006. Livelihood adaptation to climate variability and change in drought-prone areas of Bangladesh. Institutions for Rural Development No. 5. FAO

World Bank. 2000. Bangladesh: Climate Change and Sustainable Development. Report No. 21104-BD. Rural Development Unit, South Asia Region, The World Bank.

World Bank. 2008. World Development Indicators. Economy. Washington, DC: World Bank.

Page 282: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

282

Climate Change and Agricultural Development: Case Study in Cambodia

Introduction

In comparison to other Southeast Asia countries, Cambodia is not highly exposed to climate hazards. Extreme events such as typhoons or severe storms are rare because the country is protected by surrounding mountain ranges (Vanna 2000). However, it is one of the most vulnerable countries to climate change as a result of its low adaptive capacity (Yusuf and Francisco 2009). About 40 percent of the population lives on less than U$1.25 a day (Bauer et al. 2008). Agriculture employs about 68 percent of the labor force and contributes 30 percent to GDP (World Bank 2008; FAO 2004). Furthermore, only about 7 percent of total cropland is irrigated (World Bank 2008). A high dependence of the economy on the agricultural sector combined with low irrigation rates make the country highly sensitive to any changes in climate.

Areas most vulnerable to climate change in the country are those that lie in the Mekong Delta and the country’s coastal zone of approximately 435 km. The coastal zone, which is vulnerable to sea-level rise, plays an important role in agriculture, fishery, tourism and sea transport (Vanna 2000). According to the Ministry of Environment (2005), a one meter sea level rise would submerge 56 percent of Koh Kong City, for example.

Although the country suffers less from natural disasters such as typhoons than other surrounding countries, it experiences frequent floods and droughts. Such climate extreme events have been recognized as main causes of poverty in Cambodia (Ministry of Environment 2005a). The country experiences two types of flooding events, one related to overflow of the Mekong and Tonle Sap Rivers and the second as a result of extreme rainfall (Ministry of Environment 2005a). Floods and droughts occur on a yearly basis in rice producing provinces in Cambodia causing considerable economic and social losses. In 2000, for instance, severe floods killed several hundred people and caused US$150 million in damages to crops and infrastructure, while a severe drought in 2002 affected more than 2 million people and destroyed more than 100,000 ha of paddy fields (Ministry of Environment 2005a).

Land degradation is likely to intensify the impacts of a changing climate in the Cambodia coastline. The removal of coastal vegetation, agricultural development and other human activities have increased sensitivity to erosion (Vanna 2000). Underground water salinisation and seawater intrusion are common problems in coastal lowland areas used for agriculture and have been reported in the Koh Kong, Sihanoukville, Kep and Kampot Provinces (Ministry of Environment 2005b).

Impacts

Temperature is expected to increase in Cambodia by 2100 in the range of 1.35oC to 2.5oC depending on model and scenario. Rainfall also varies according to model, scenario and location. By 2100 and under the A2 scenario, annual rainfall is expected to increase between 3 and 35 percent depending on location (Ministry of Environment 2001).

Losses in rice production in Cambodia are mostly associated with flood events (more than 70 percent) and droughts (about 20 percent), followed by pests and diseases (10 percent). Under climate change it is expected that wet season rice yields will increase and dry season rice yields will remain the same or decrease (Ministry of Environment 2001). However, if floods and droughts become more frequent and intense, results will be more unpredictable.

Page 283: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

283

Coping/Adaptation Policies

A survey of 684 households in 17 provinces shows how rural communities are coping with frequent floods and droughts (Ministry of Environment 2005b). Almost 20 percent of the surveyed households said that they did not make any preparations for past floods and 17 percent just planted their crops as usual. Traditional coping measures that are in place include elevated enclosures for livestock, increasing the household’s foodstock, increasing feedstock for animals, and preparing boats. Few people move to safer place in anticipation of floods. During drought events, 24 percent of villagers organised religious ceremonies to ask for rain only. About 16 percent just planted crops as usual and 17 percent reduced water consumption. Around 12 percent of people constructed wells and 11 percent pumped water (Ministry of Environment 2005b). Vulnerability to drought and floods is increased by the fact that 95 percent of surveyed households stated farming as their main source of income and 92 percent mention rice farming as the main income.

According to the Ministry of Environment (2005), in order to achieve food security, total rice planted area should reach 2.5 million ha with a productivity of about 2.2 tons per ha, requiring significant agricultural productivity improvement. The Government of Cambodia expects to considerably increase the irrigated areas of the country as a way to increase food security and better cope with climate change (Ministry of Environment 2005a). The Ministry of Water Resource and Meteorology goal is to implement irrigation projects including rehabilitation of pumping stations and water pumps. Until 2003, 315 irrigation projects had been implemented covering 153,149 ha of paddy rice (Ministry of Environment 2005a).

In broad lines, the National Action Plan on Climate Change recommends the following adaptation and mitigation strategies for the agricultural sector (Ministry of Environment 2005a).

Promotion of least GHG emission agricultural practices

Improvement of consumption of non-rice staple foods (crop diversification)

Expansion of the best available rice planting systems for suitable land areas

The Ministry of Environment (2005a) also mentions that with the exception of the National Adaptation Program of Action to Climate Change (NAPA) and the National Biodiversity Strategy and Action Plan, there is no government policy or legislation in Cambodia that mentions the need for adaptation to climate change in coastal areas. According to Vanna (2000), few studies have been done to assess the impacts of climate change on Cambodia’s coastal zone and possible adaptation strategies. There is also very limited knowledge of the local population and authorities on impacts of climate change.

Cambodia heavily relies on external funding for implementation of programmes. In water resources, for instance, in the period of 2004-2006, over 85 percent of the funding came from external sources (Ministry of Environment 2005a). However, if Cambodia wishes to optimize the use of funding for climate change adaptation policies, policy makers and other stakeholders need to improve their understanding of climate change impacts and adaptation strategies. Information needs also to be disseminated to rural communities which are the most affected by droughts and floods (Ministry of Environment 2005a).

Currently, the weather station network in Cambodia is very limited. Most of it was destroyed by the war. There are only 12 rainfall stations in the whole country. Air temperature is

Page 284: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

284

recorded only in six locations. Climate forecasting in the country is very limited (Ministry of Environment 2005a).

A household survey shows that villagers find the interpretation of flood data – water level data - posted in public places very hard to understand and there is no clear information about when flooding will occur. The vast majority of surveyors state that they never received any form of early drought warning (Ministry of Environment 2005b). Furthermore, 45 percent of interviewees stated that they did not receive any sort of assistance after the occurrence of natural disasters while 30 percent received assistance from the Cambodian Red Cross.

Agricultural extension services in Cambodia also seem to be limited. Average wet season and dry season rice yields in rural Cambodia are less than 2.0 and 3.3 t/ha respectively. However, the theoretical yields range from 3.5 to 6.0 t/ha—yields easily achieved on the Viet Nam side of the Mekong Delta—which might be an indication of limited transfer of technology and knowledge to farmers and the need for access to more crop inputs (Ministry of Environment 2005a). Moreover, as Cambodia has poor quality soil, better land suitability assessment should be conducted to replace unsuitable crops with more suitable ones (Kirby and Mainuddin 2009).

Cambodia lacks basic capabilities to deal with climate variability and recurrent drought and floods such as early warning systems, research and extension services that aim to make agricultural more resilient to climate extremes and social networks that provide assistance to rural communities. As a result, even small changes in climate can greatly increase the vulnerability of poor rural households and their food insecurity.

References

--------. 2005a. Analysis of policies to address climate change impacts in Cambodia. Phnom Penh: Ministry of Environment.

--------. 2005b. Vulnerability and adaptation ot climate hazards and to climate change: a survey of rural Cambodian households. Phnom Penh: Ministry of Environment.

Bauer, A., R. Hasan, R. Magsombol, and G. Wan. 2008. The World Bank's New Poverty Data: Implications for the Asian Development Bank. ADB Sustainable Development Working Paper Series No. 2. Manila, Philippines: Asian Development Bank

FAO. 2004. FAOSTAT - Agricultural Employment. Rome: FAO.

Kirby, M. and Mainuddin, M. 2009. Water and agricultural productivity in the Lower Mekong Basin: trends and future prospects. Water International,34:1,134 — 143.

Ministry of Environment, 2001. Vulnerability and adaptation assessment to climate change in Cambodia. Phnom Penh: Ministry of Environment.

Vanna, P. 2000. Potential impacts of climate change on Cambodia coastal zone. Report from field survey at Kampong Som Province, 28 Nov - 01 DEc, 2000.: Climate Change National Technical Committee

World Bank. 2008. World Development Indicators. Washington, DC: World Bank.

Yusuf, A.A., and H. Francisco. 2009. Climate change vulnerability mapping for Southeast Asia. South Bridge Court, Singapore: EEPSEA.

Page 285: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

285

Climate Change and Agricultural Development: Case Study in the Fiji Islands

Introduction

Fiji is comprised of over 300 islands. Viti Levu and Vanua Levu are the largest islands at 10,429 km2 and 5,556 km2, respectively. Viti Levu accounts for approximately 70 percent of Fiji’s population (of about 830,000 people) and has much of the land used for sugarcane farming in the country (Agrawala et al. 2003). The economy of Fiji is highly dependent on the performance of two sectors, tourism and agriculture (sugar production). However, since the 1990s, the contribution of the agricultural sector to the GDP has been steadily declining (from 19 percent in 1996 to 12.5 percent in 2006) while the service sector has been increasing the GDP (from 58.4 percent in 1996 to 65.9 percent in 2006) (ADB 2008). In 1999, agriculture accounted for 43 percent of Fiji’s foreign exchange earnings. Currently the tourism sector is the largest foreign exchange earner (PICCAP 2005).

Despite the growing tourism sector, Fiji is still highly dependent on exports from sugar. The agricultural sector is still the main employer in Fiji. In 2007, while the service sector employed 37,400 people, the agricultural sector employed 119,800 people (ADB 2008). The sugar industry alone employs about one fifth of the national workforce (Kenny et al. 2000).

About 24 percent of the arable land of the country is cultivated with sugarcane, 23 percent with coconut, and the remaining 53 percent with other crops. The three main types of agricultural activity are subsistence farming, semi-commercial farming, and plantation farming. Subsistence farming contributes to over 30 percent of the agricultural GDP and is based up on staple root crops (e.g., dalo, cassava, yams, and sweet potato) and tree crops (e.g., coconuts, bananas, breadfruit, mangoes, and other fruit trees), rice, and vegetables. Semi-commercial farming products are supplied to urban areas and exported. The plantation farming is mainly comprised of sugarcane and coconut (Kenny et al. 2000).

The Fiji Islands are very prone to El Niño events, a major cause of droughts in the country, and to cyclones that constantly bring floods and landslide. Such events are main causes of economic losses and infrastructure damage (PICCAP 2005). In addition, over 90 percent of the population and the majority of services, agricultural production, and infrastructure are on the coasts (PICCAP 2005), which makes the country highly vulnerable to sea-level rise and climate extreme events. Only 16 percent of the countries’ total land area is suitable for farming, which is located mainly along coastal plains, river deltas, and valleys of the two main islands, Viti Levu and Vanua Levu (PICCAP 2005). Furthermore, about 25 percent of Fiji’s population lives below the poverty line (Agrawala et al. 2003), which might indicate that the population has a low capacity to adapt to climate change.

Impacts of Climate Change on Agriculture and Food Security

Fiji’s contribution to carbon dioxide emissions is considered insignificant due to its small size and development status (Kenny et al. 2000; UNDP 2008). On the other hand, being a small country, Fiji is likely to be disproportionally impacted by climate change, sea-level rise, and more intense climate extreme events.

There is a limited understanding of future changes in the frequency and intensity of climate extreme events. However, the IPCC 4th Assessment Report suggests that such events are projected to increase in intensity in many regions of the world (Meehl et al. 2007). There is also moderate confidence that in the Pacific Islands region, by 2050, the intensity of cyclones will

Page 286: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

286

increase by as much as 20 percent and that El Niño events will become more common (World Bank 2000).

Past losses associated with climate extreme events show that projected increases in climate variability as a result of climate change will have a catastrophic effect on the population of Fiji. The most severe drought Fiji has experienced in 1997-98 resulted to losses in sugar production and thus generated an economic crisis where the country has a negative growth of 4 percent. With the drought condition ended and crop production was back to normal, the economy recovered (PICCAP 2005). After the catastrophic drought where over 263.000 people were affected, flooding associated with La Niña broke in the western region (Kenny et al. 2000). During the current decade, natural disasters have continued to adversely impact Fiji. In February 2007, a flood event caused economic damages of US$30 million. The same amount of economic damages was caused by a storm in 2003 affecting 30,000 people (EM-DAT22).

Sugarcane, in particular, is negatively affected by droughts. Kenny et al. (2000) discussed the present expectation in Fiji is for an annual sugarcane production is around 4 million tons. However, that amount of production was from 1978 to 2007. As Figure 1 indicates, during the El Niño years (1978, 1983, 1987, 1992, 1998, and 2003) sugarcane production was significantly below of average, particularly in 1983 and 1998. The impact of the 1997-98 drought was greatest in marginal sloping lands and in sandy, coastal soils, revealing the vulnerability of those who do not have access to flat land. According to Kenny et al. (2000), 75 percent of the land area used for sugarcane production is considered marginal for rainfed production. Floods, heavy rainfall, and cyclones can also adversely impact sugarcane production causing waterlogging and soil erosion in marginal lands.

According to the World Bank’s estimates (2000), there is moderate confidence that as a result of temperature and rainfall changes and El Niño, agricultural economic damages will increase in Viti Levu, resulting in losses in sugarcane, yams, taro, and cassava. The impact of climate change on agriculture in Viti Levu is estimated to cost about US$14 million a year by 2050 (Table 1). This estimate could be much higher depending on the severity of extreme events.

In the event of climate changes in Fiji, the most significant economic damage would be on sugarcane. However, losses of traditional crops, such as yams and taro, could have a substantial effect on subsistence economies (World Bank 2000). In Viti Levu, dalo and yam production would be little affected by projected changes in average climate conditions. However, during El Niño years, dalo yield could be reduced by 30-40 percent of current levels by 2050. On the other hand, yam production might remain the same or increase during El Niño events and decline by around 50 percent during La Niña wetter events. Cassava production is expected to decline as a result of changes in average climate conditions (5 to -9 percent by 2050) and during La Niña events. Also in Viti Levu, sea level rise causing shoreline to retreat could cause land losses to erosion with damages estimated in US$5.8-7.3 million in 2050 and US$8.2-18.6 million by 2100 (World Bank 2000).

Adaptation Strategies

Many adaptation strategies have been implemented in the Pacific Island countries. Several projects have been done to protect coastal dwellers and structures against coastal erosion and storm surges through the construction of seawalls. Some of these strategies have, however, proven to be unsuccessful indicating a need for better evaluation of adaptation options. In the

22 EM-DAT – Emergency Events Database (http://www.emdat.be/).

Page 287: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

287

Qoma Islands, for instance, frequent inundation further downstream after the construction of a sea wall upstream has been reported (Mataki, Koshy, and Nair 2006).

Better evaluation of strategy options is also essential for the agricultural sector. Kenny et al. (2000) suggest some adaptation measures to increase the resiliency of the agricultural sector in Fiji, which are listed below:

Cassava

If tropical cyclones get more intense as predicted, the trend toward increasing the cultivation of cassava might lead to greater damages than if traditional root crops are maintained. Understanding the reasons for such trend and promoting sustainable agriculture would be an important first step.

Sugarcane

Marginal sloping and coastal lands should not be used for sugarcane production. Alternative crops combined with more sustainable land-use practices are an alternative. The viability of irrigation in the better land should be assessed, which would increase sugarcane yields and reduce harvest areas. Figure 2 indicates that while the harvest area has increased, sugarcane yields have remained low.

Root crops

As opposed to cassava, root crops might have a high adaptive capacity. Some adaptation measures could be to identify most suitable land areas for root crops, breed more drought tolerant dalo varieties, and to enhance the yam breeding programme.

Other adaptation options include climate-proofing farming systems, promotion of sustainable production systems, promotion of land use planning and improved seasonal forecasting (World Bank 2000). Kenny et al. (2000) also highlights that there is high uncertainty with the land-lease situation in Fiji. As a result, Fijian landowners have little incentive to diversify crops and to implement soil and water conservation measures. Crop diversification would be an option if such uncertainty regarding land lease were reduced (Kenny et al. 2000).

As for most of the Pacific Island countries, there is a serious lack of current data for the Fiji Islands related to socioeconomic, climate, land degradation, water, and weather forecast indicators. Such limitation makes the assessment of the country’s vulnerability to climate change and potential adaptation strategies more challenging.

References

ADB. 2008. Key indicators for Asia and the Pacific 2008 - Fiji Islands. Tunis: ADB. Available at <http://www.adb.org/Documents/Books/Key_Indicators/2008/pdf/fij.pdf> . Accessed April 14, 2009.

Agrawala, S., T. Ota, J. Risbey, M. Hagenstad, J. Smith, M. van Aalst, K. Koshy, and B. Prasad. 2003. Development and climate change in Fiji: Focus on coastal mangrove. No. COM/ENV/EPOC/DCD/DAC(2003)4/FINAL. Paris: OECD.

FAO (Food and Agriculture Organization of the United Nations). 2007. FAOSTAT/PRODSTAT module. <http://faostat.fao.org/site/526/default.aspx>. Accessed April 18, 2009.

Page 288: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

288

Kenny, G., J. Bhusan, R. Ogoshi, and I. Ratukalou. 2000. Agriculture. In Climate change vulnerability and adaptation assessment for Fiji . IGCI Technical Report . Ed. J. Feresi, G. Kenny, N. de Wet, L. Limalevu, J. Bhusan, and I. Ratukalou. Washington, DC: The World Bank.

Mataki, M., K. Koshy, and V. Nair. 2006. Implementing climate change adaptation in the Pacific Islands: Adapting to present climate variability and extreme weather events in Navua (Fiji). AIACC Working Paper No. 34. Washington, DC: International START Secretariat.

Meehl, G. A., T. F. Stocker, W. D. Collins, P. Friedlingstein, A. T. Gaye, J. M. Gregory, A. Kitoh, R. Knutti, J. M. Murphy, A. Noda, S. C. B. Raper, I. G. Watterson, A. J. Weaver, and Z. Zhao. 2007. Global climate projections. In Climate change 2007: The physical science basis. Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change. Ed. S. Solomon, Qin.D., M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor, and H. L. Miller. Cambridge: Cambridge University Press.

PICCAP. 2005. Fiji's First national communication under the framework convention on climate change. Suva, Fiji islands: Department of Environment.

UNDP. 2008. Human Development Report 2007/2008 Country Fact Sheets - Fiji. New York: UNDP. Available at <http://hdr.undp.org/en/statistics/> . Accessed March 7, 2009.

World Bank. 2000. Cities, sea, and storms. Managing change in Pacific Islands economies. Volume IV Adapting to climate change. Washington, DC: Papua New Guinea and Pacific Island Country Unit, the World Bank.

Page 289: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

289

Table 1.Estimated economic impact of climate change on agriculture in Viti Levu, Fiji, 2050

Source: World Bank 2000

Page 290: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

290

Figure 1: Sugar cane production in Fiji (tons)

Source: FAO 2007

0

500,000

1,000,000

1,500,000

2,000,000

2,500,000

3,000,000

3,500,000

4,000,000

4,500,000

5,000,000

1978

1979

1980

1981

1982

1983

1984

1985

1986

1987

1988

1989

1990

1991

1992

1993

1994

1995

1996

1997

1998

1999

2000

2001

2002

2003

2004

2005

2006

2007

Page 291: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

291

Figure 2: Sugarcane yield and harvested area in Fiji

Source: FAO 2007.

0

10000

20000

30000

40000

50000

60000

70000

80000

1978

1979

1980

1981

1982

1983

1984

1985

1986

1987

1988

1989

1990

1991

1992

1993

1994

1995

1996

1997

1998

1999

2000

2001

2002

2003

2004

2005

2006

2007

Area Harvested (Ha) Yield (Kg/Ha)

Page 292: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

292

Climate Change and Agricultural Development: Case Study in THE Maldives

Introduction

The Maldives archipelago comprises 25 geographic atolls. The country’s atolls contain a total of 1,190 islands, of which 358 currently are being used mainly for human settlements, infrastructure and economic activities (The Government of Maldives 2006). All the Maldives islands are low lying and none exceed the elevation of three meters. More than 80 percent of the land area is less than one meter above mean sea level (MHTE 2001). Therefore, Maldives is one of the most vulnerable countries in the world to sea-level rise. Besides being among the five countries in the world comprised entirely of low-lying atolls, the country also has a high population density (909 people/km2), which means that large numbers of people are potentially exposed to single events (Barnett and Adger 2003).

The main economic activity of the country is tourism, which accounts for one third of the GDP. Fisheries, the largest contributor to exports, account for 7 percent of the GDP, while agriculture plays a minor role in the economy, accounting for only 2.8 percent of the GDP (The Government of Maldives 2006). However, the agricultural sector is still important for food security, nutrition and employment opportunities for the country’s communities as 75 percent of the inhabited islands have some degree of agricultural activity (MPND 2007).

Thirty-two uninhabited islands of the country are leased for a 21-year period for commercial farming. In those islands, in 2004, watermelon accounted for 33.6 percent of total production, followed by papaya (26 percent) and coconut (11.5 percent) (MPND 2007). In the islands’ subsistence agriculture, the main crops are banana, watermelon, cucumber, taro, coconut, breadfruit, mango, sweet potato, pumpkin, papaya, luffa, cabbage and brinjal (The Government of Maldives 2006). Coconut is the most common plantation crop in all the atolls and the most popular home garden tree (MHTE 2001).

The low participation of the agricultural sector in the Maldivian economy is in part a result of the bad quality of the land. Suitable land for agriculture in the Maldives is estimated to be less than 30 m2; as a result, 90 percent of the country’s food demand is met from imports (MPND 2007). Except for tuna and coconut, the Maldives imports all food items. Rice, the main staple food, is mostly imported from South and Southeast Asia (MHTE 2001). Annually, 17 million kilograms of rice, the same quantity of flour and 10 million kilograms of sugar are imported every year (The Government of Maldives 2006). On the other hand, fish production in the Maldives tripled since 1970 as a result of vessel mechanization and the introduction of fish freezing and canning. Skipjack tuna constitutes around four-fifths of total production (MPND 2007). As a result, fisheries account for more than 80 percent of total agricultural export and around 40 percent of total merchandise exports (FAO 2005).

With its high dependency on food imports, the impact of climate change on food security in Maldives is not tied solely to its agricultural sector. Overall, Maldives’ economy is highly vulnerable to climate change. Climate change, and sea level rise in particular, poses a threat to the tourism and fisheries sector, the largest of the Maldives’ economy. The country’s ability to import staple food and therefore guarantee food security of the population depends on revenues coming from those sectors. It also depends on how climate change will affect agricultural production in countries that export to Maldives.

Page 293: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

293

Impact of Climate Change on Agriculture and Food Security

The Maldives contribute 0.001 percent of global greenhouse gas emissions, with the energy sector being the main contributor. Agricultural emissions are insignificant (MHTE 2001). On the other hand, global warming and more frequent climate extreme events will severely affect the Maldives’ agriculture industry, intensifying the impacts of current stressors such as the country’s poor soil, limited land available for cultivation and water scarcity (The Government of Maldives 2006). Those most adversely affected by climate change are likely to be the vulnerable and poor citizens of the country who work in the agriculture, fishing and local manufacturing sectors.

The drastic impact of extreme events on agriculture was seen with the 2004 tsunami. Among affected countries, Maldives’ agricultural sector was one of the most severely hit. Crops were severely damaged by salinization of soil and salt water intrusion into the aquifers. Field crops were damaged in 2,103 farms. In 11.678 homesteads, backyard crops and agricultural tools were destroyed, and 700,000 fruit trees were damaged in the inhabited islands (MPND 2007). Extreme climate events also pose a risk to food security because of the country’s limited long-term and emergency food storage. Furthermore, food distribution to the islands, which is basically done by boat from the capital, is threatened by extreme events, increasing the risk of food insecurity (The Government of Maldives 2006).

Sea level rise and wave-induced flooding are likely to increase saltwater intrusion into the freshwater lens, threatening the little agriculture that exists in Maldives (MHTE 2001; The Government of Maldives 2006). Problems with freshwater aquifers have already damaged fruit trees in cities like Male, where most of the mango trees have died. Besides mango, other deeper-rooted trees with low salt tolerance such as banana and breadfruit are likely to be adversely affected by saltwater intrusion of groundwater. Taro is another crop that is likely to be negatively affected by the rising sea level. Taro is grown in taro pits dug in and near the wetlands, about 40 cm above mean sea level. Currently, these pits are already below the highest tidal levels (MHTE 2001). Climate change also threatens to drive the tuna stock – the main fishery of Maldives – to more favorable temperatures (UNEP 2001).

The impact of sea level rise on groundwater has potentially serious consequences. In Maldives, groundwater is a scarce resource. The freshwater aquifer that lies beneath the country’s islands is a shallow lens located 1 to 1.5 m below the surface and only a few meters thick. Rainwater is the principal source of drinking water for 90 percent of the atoll households. In Male, the country’s capital, 100 percent of the population has access to piped desalinated water. After the tsunami, 38 islands started using desalination plants (The Government of Maldives 2006).

Adaptation Strategies

Given the severe impact of rising sea levels on Maldives, much of its mitigation efforts revolve around increasing global awareness of the impacts of climate change on small islands countries. In international forums, Maldives have sought to bring attention to the vulnerability of low-lying small island developing states to climate change (Ministry of Environment & Construction 2004).

According to the Government of Maldives, because the country is composed of tiny islands where physical space is scarce and the land is flat and low lying, not all adaptation measures to cope with sea level rise are viable. For instance, retreat, raising of the land and the use of building setbacks may not be viable as that would involve abandoning the coastal zone

Page 294: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

294

and shifting the associated ecosystems inland (Ministry of Environment & Construction 2004). Protection of the coastal zones through application of solid structures such as seawall might be a more realistic option along well-developed coasts. In Male, the Japanese Government supported the construction of a seawall. However, similar structures would be needed in other inhabited islands, which will require financial and technical assistance from international aid agencies and donors (Ministry of Environment & Construction 2004).

The Maldives’ government has prioritized the protection of its vital infrastructure including the airports, which are essential for the food security of the country as most of the consumed food is imported. Other efforts from the government have been to strengthen flood defenses with the development of innovative tetrapods in Male; to raise public awareness and promote behavioral change from Maldivians and tourists; and the development of Hulhumale (an artificial island) to reduce overcrowding in Male. Finally, the government is considering the concept of a safe island zone, where vulnerable communities would be identified and relocated to places where they could better build their livelihoods (Latheef 2007).

Several adaptation measures for all sectors vulnerable to climate change were proposed by the Government of Maldives in its National Adaptation Programme of Action (NAPA). According to NAPA, increasing local food production is a key adaptation measure to increase the country’s food security. A new project was designed to increase the capacity of farmers, local communities and institutions to cope with climate change by enhancing their knowledge, access to technology, best practices, marketing and pest control. The total cost of the project was estimated to be US$825,000 (The Government of Maldives 2006).

Important issues, however, obstruct the development of successful adaptation strategies (Ministry of Environment and Construction 2004):

Lack of capacity to adapt both financially and technically in all important sectors makes

it more difficult to successfully respond to the challenges of climate change. Human resource development and institutional strengthening are urgent requirements.

There is insufficient research and systemic observation. Maldives has only three stations that measure sea level and five meteorological stations that measure only the basic parameters required for general weather forecasting. More stations are needed to study rainfall patterns, as they vary greatly at different locations in Maldives. Also, more research is needed to understand issues and adaptation measures related to beach erosion and growth patterns of coral reefs.

However, if the sea level continues to rise, the people from Maldives will have more urgent worries. Sea level rise might threaten to drown the archipelago, directly affecting their sovereign territory. In 2008, President-elect Mohammed Nasheed mentioned an interest among the islanders to buy a new homeland (Bogardi and Warner 2008). If that in fact happens, the main adaptation strategy will be to think how to relocate and resettle the entire population of Maldives.

References

Barnett, J. and W. Adger. 2003. Climate dangers and atoll countries. Climatic Change 61(3): 321-337.

Page 295: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

295

Bogardi, J. and K. Warner. 2008. Here comes the flood. Nature reports climate change. Published online 11 December 2008. doi.10.1038/climate.2008.138

FAO. 2005. Small Island Developing States. Agricultural Production and Trade, Preferences and Policy. FAO Commodities and Trade Technical Paper No. 7.

Latheef, M. 2007. Maldives. Statement by His Excellency Dr. Mohamed Latheef, Permanent Representative of the Republic of Maldives to the United Nations at the Informal Thematic Debate of the General Assembly on the Theme "Climate Change as a Global Challenge." United Nations Headquarters, New York.

MHTE. 2001. First National Communication of Maldives to the UNFCCC. Male: Ministry of Home Affairs, Housing and Environment.

Ministry of Environment & Construction. 2004. State of the Environment. Maldives. Male: Ministry of Environment & Construction.

MPND. 2007. Seventh National Development Plan. 2006-2010. Male: Ministry of Planning and National Development. Republic of Maldives.

The Government of Maldives. 2006. National Adaptation Programme of Action (NAPA). Male: Ministry of Environment, Energy and Water.

UNEP. 2001. State of Environment Maldives 2002. Bangkok: UNEP/RRC.AP.

Page 296: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

296

Climate Change and Agricultural Development: Case Study in Pakistan

Introduction

Land degradation, frequent climate extreme events, and a high dependency on agriculture and irrigation make the Pakistani rural population highly vulnerable to climate change. About one fourth (20 million hectares) of Pakistan’s total land area is arable, of which 16 million hectares (ha) have irrigated agriculture (O'Brien 2000). Agriculture contributes to 20 percent of the Pakistani GDP (livestock alone accounts for 11 percent of the GDP) and employs about 43 percent of the total labor force (World Bank 2008). About 81 percent of farmers own areas that are less than 5 ha, constituting 39 percent of total cultivated area. The main cropping patterns in practice are rice-wheat, cotton-wheat, sugarcane-wheat, and maize-wheat (Ministry of Environment 2003).

Agricultural lands in Pakistan already experience severe degradation, which makes the country more vulnerable to climate change. About 22 percent of the country’s land suffers from severe land degradation (water erosion, wind erosion, and chemical deterioration), 20 percent from salinity, 24 percent from shallowness, and 13 percent from erosion risk (Bot et al. 2000). Possibilities of land extensification are, however, limited. There is very little additional land available for agriculture, and more intensive land use is limited by increasing water scarcity. As a result of a rising population, falling water flows, and erosion in the storage capacity, per capita water availability is declining in Pakistan and it is currently only marginally above the threshold level of water scarcity of 1,000 cubic meters (Mustafa 2004). Therefore, Pakistan has moved from being a water-affluent country towards being a water-scarce country (O'Brien 2000).

Pakistan is considered a high human vulnerability country to natural hazards, with areas that are flood or drought hotspots (Ehrhart et al. 2008). Besides floods and droughts, it is also often hit by extreme temperatures, landslides, earthquakes, and windstorms. In 2005, for instance, over 73,000 people died and 3 million people were affected by the South Asian earthquake. That disaster caused the highest death toll in that year (Sanker, Nakano, and Shiomi 2007). In 2007, Pakistan was the fifth most affected country in the world by extreme weather events with associated damages of 0.62 percent of its GDP. In the same year, it was also the fourth country with the highest death toll--flooding events were the main responsible for death tolls and economic losses (Harmeling 2008).

All the aspects mentioned above contribute to make agriculture and rural communities in Pakistan vulnerable to climate change. Furthermore, high poverty rates—almost a fourth of the country (22.59 percent) lives under the poverty line of US$1.25 a day (Bauer et al. 2008)—indicate that adaptive capacity of the country might be low, particularly, in rural areas where poverty rates are higher.

Impacts of Climate Change on Agriculture and Food Security

In Pakistan, temperatures have increased since the 1970s, increasing the heat on plants that are already at the margin of stress. Increases in temperature might particularly affect wheat, cotton, mango, and sugarcane, as the prevailing maximum temperature is more than 10ºC higher than the optimal range (Ministry of Environment 2003). Average temperature is projected to increase from 19.99 ºC (average of 1960-1990) to 22.20 ºC by 2080 (IFPRI estimate). However, the largest impacts of climate change on agriculture and other sectors are likely to be a result of the increased variability of the monsoon (Reid, Simms, and Johnson 2007). Scenarios project that even if precipitation increases, the expected outcome might not be entirely positive. A

Page 297: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

297

combination of increased river flows and flood plain pressures could make flood hazards one of the most serious problems associated with climate change (O'Brien 2000). In the Indus River basin total annual run-off from the upper basin is likely to increase by 11 to 16 percent possibly aggravating problems of flooding, waterlogging, and salinity in the basin (Kelkar and Bhadwal 2007). The lack of capacity to storage water in the Indus River Basin means that increased precipitation would not necessarily lead to improved water use (O'Brien 2000). Overall, floods in agricultural areas are likely to lead to salinity, chemical contamination, or soil erosion, affecting food security of the country (Reid, Simms, and Johnson 2007).

Coping/Adaptation Strategies

Under a scenario of overall water surplus, water shortages might still occur in local areas of the highly productive Punjab rice-wheat zone and in the unglaciated valleys of the upper basin (Kelkar and Bhadwal 2007). Any change in precipitation levels can exacerbate the vulnerability of rural communities, particularly the ones living in the deserts that depend on rainwater as a primary source of fresh water. A survey in Bahawalpur showed that poor farmers had limited ways to cope with droughts and were in need of adaptation measures (Mirza 2007). The survey found that during drought events, the main coping mechanism of farmers was to sell the livestock followed by the sale of belongings, increasing even more their vulnerability. Furthermore, farmers did not take any significant precautions to preserve seed and did not apply any seed treatment. The surveyed farmers were unaware of the modern techniques of crop husbandry to cater to the problems of brackish water.

On the other hand, according to the Pakistani government, some farmers have already adopted several measures to cope with environmental constraints. Shortages of water and changes in plantation and irrigation methods have made farmers grow onions, potatoes, tomatoes, cauliflower, and cabbage on ridges using furrow irrigation. Some progressive farmers have already started to reduce the area under sugarcane , as they are expecting a rise in temperatures and a decrease in rainfall that would increase the net irrigation water requirement of crops. Farmers are considering replacing sugarcane (damaged during droughts) with sugar beet , as the sugar industry of the country has showed interest in using beet for sugar production. Experiments in the Sindh province with cultivation of sugar beet have shown to be promising. Agricultural scientists have also started crop-breeding programs to develop drought and flood tolerant crops in a response of extreme variability due to monsoons (Ministry of Environment 2003).

As mentioned above, such actions aimed at adapting agriculture to climate change were done by progressive farmers who most likely had access to information and financial resources. Poor farmers who do not have access to information, assistance, and financial resources are the most vulnerable to increased water scarcity, land degradation, and higher incidence of floods and droughts. Hence, poor farmers are the most in need of adaptation measures. In a country such as Pakistan that faces water scarcity, extreme events, and heat stressed crops, one of the most important adaptation strategies to better cope with climate change and decrease rural communities’ vulnerability is the conservation of water and the development of sustainable water projects (O'Brien 2000; Ministry of Environment 2003).

A project by Oxfam shows that adaptation strategies can in fact reduce the vulnerability of communities to climate events. In Southern Punjab, Oxfam started a project to reduce vulnerability of communities to floods. The organization used several participatory approaches to create community-based organizations with the goal of making communities better prepared to

Page 298: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

298

deal with climate extreme events. Main activities included: building community awareness and forming community-based organizations; networking with the government; doing small-scale disaster mitigation work (building emergency shelters and raising homesteads), developing community-based early warning systems, and promoting training in social forestry techniques and soil and water testing.

The benefits of such community-based adaptation strategies were seen in the summer of 2006, when early warning committees in the villages of the Mizafargarh District in south Punjab got timely information about the arrival of floods. As a result, the communities were able to take proactive action by building a 22-kilometer bund (embankment) around five villages. The 2006 flood was more intense than the 2005, however, the bund resulted in smaller losses of agricultural land in 2006 (Reid, Simms, and Johnson 2007).

Still, much has to be done at all levels. Some studies suggest potential adaptation strategies in the water and agriculture sector, which are listed below (O'Brien 2000; Mirza 2007; Reid, Simms, and Johnson 2007; Ministry of Environment 2003; Shaukat 2008):

Long-term strategies need to consider more efficient and sustainable water use for

agricultural purposes. Low water demanding varieties should be prioritized as opposed to water intensive crops. Nonwater management options such as tillage, precision planting, plant nutrition, drainage, and salinity management should also be prioritized.

Sustainable land management is essential in Pakistan considering the already high degree of soil degradation in the country. More sustainable land use could be done by using organic composts, integrated pest management, and effective microbes for enrichment of organic materials. Integrated management leads to ecological improvements and helps rural communities to develop sustainable livelihoods.

Small dams could contribute to increase Pakistan’s water storage capacity. The most ideal location would be the Jhelum River because of the relatively small investments.

Village ponds have been degraded due to silting and weed growth. The ponds can help store rainwater that falls in the monsoon season. Therefore, they should be regularly cleaned, and existing depressions should be maintained. More ditches and depressions should be found for increased water storage.

More responsibilities for operations should be given to farmers. Improvements in communication and mass education could facilitate improvements in the water distribution system allowing farmers to have greater responsibilities for operations in an attempt to increase cropping intensities (considering the low availability of additional arable land in Pakistan).

Livestock low nitrogen feeds and controlled fermentation through diet management should be encouraged.

Conversion technologies should be introduced and controlled burning should take place. Adaptation projects need to have a component aimed at increasing knowledge and

information of farmers. A survey in Bahawalpur found that significant investments occurred in farm water management. However, the largest majority of farmers did not buy into the effectiveness of on-farm water-user associations promoted by various international development organizations.

The government should take a more active role. The P government’s draft of the National Water Policy underscores the need for assessing the impact of climate change and monitoring

Page 299: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

299

and factoring the impact into future water development strategies. Farmers from Bahawalpur say that the most needed measure from the government is the provision of canal water.

Adaptation measures aimed at reducing poor farmers’ vulnerabilities to global warming need to consider multiple factors when addressing the impact of climate change. Low level of economic and social development, high levels of land degradation and water scarcity, and inadequate infrastructure, among others, increase the vulnerability to climate change of poor farmers, who are highly dependent on the natural resource base. Furthermore, farmers need to be prepared not only for a change in mean temperature and rainfall but also to the incidence of more frequent and intense floods and droughts events. An essential component in any successful adaptation strategy will be to increase farmers’ access to information.

References

Bauer, A., R. Hasan, R. Magsombol, and G. Wan. 2008. The World Bank's New Poverty Data: Implications for the Asian Development Bank. ADB Sustainable Development Working Paper Series No. 2. Manila, Philippines: Asian Development Bank.

Bot, A. J., F. O. Nachtergaele, and A. Young. 2000. Land resource potential and contraints at regional and country level. World Soil Resources Reports No. 90. Rome: FAO.

Ehrhart, C., A. Thow, M. de Blois, and A. Warhust. 2008. Humanitarian implications of climate change. Mapping emerging trends and risk hotspots. Chatelaine, Geneva: CARE International.

Harmeling, S. 2008. Global climate risk index 2009. Weather-related loss events and their impacts on countries in 2007 and in a long-term comparison. Bonn: Germanwatch.

Kelkar, U., and S. Bhadwal. 2007. South Asian regional study on climate change impacts and adaptation: Implications for human development. Occasional Paper No. 2007/27. New York: UNDP.

Mirza, M. 2007. Climate change, adaptation and adaptive governance in water sector in South Asia. Toronto: Environment Canada.

Mustafa, I. 2004. Water Shortage in Pakistan. Pakistan Water Gateway Available at <http://www.waterinfo.net.pk/docs.htm> . Accessed January 29, 2009.

O'Brien, K. 2000. Developing strategies for climate change: The UNEP country studies on climate change impacts and adaptations assessment. CICERO Report No. 2000:2. Oslo: CICERO, UNEP.

Pakistan, Ministry of Environment. 2003. Pakistan's initial national communication on climate change . Islamabad: Ministry of Environment.

Reid, H., A. Simms, and V. Johnson. 2007. Up in smoke? Asia and the Pacific. The threat from climate change to human development and the environment. London: NEF.

Sanker, S., H. Nakano, and Y. Shiomi. 2007. Natural disasters data book—2006. An analytical overview . Kobe, Japan: ADRC.

Shaukat, A. 2008. Climate change: The Impact on biodiversity and agriculture. Islamabad: Lead.

World Bank. 2008. World development indicators. Economy. Washington, DC: World Bank.

Page 300: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

300

Climate Change and Agricultural Development: Case Study in the Philippines

Introduction

Developing countries like the Philippines are affected by climatic variability. Agriculture contributes significantly to the economy at 14.1 percent of the estimated GDP of US$3,127 per capita in 2006 (the service sector contributed 54.2 percent, industries 31.7 percent) (ADB 2008). The country’s major agricultural commodities range from sugarcane, coconuts, rice, corn, bananas, cassavas, pineapples, and mangoes as well as livestock such as pork, eggs, beef and fish. Communities in the rural areas depend largely on agriculture for subsistence, nutrition and income, a situation similar to most countries in Asia.

Rosegrant et al. (2007) stressed that despite current agricultural technologies, changes in key drivers affecting food production such as population growth and dietary patterns combined with climatic stresses like droughts, floods, and extreme temperatures will undermine the ability of Asian farmers to produce the food needed to generate sufficient income for productive and healthy livelihoods. Evidence is emerging that the Philippines is vulnerable to these climatic stresses. Droughts and floods that occurred in the past years have substantially affected the agricultural production in the country. The ENSO23-related droughts affected the water resources, which in turn affected the agricultural sector as well as health and environment (Jose and Cruz 1999). The threats posed by climatic variability to agricultural production eventually lead to serious social and economic implications, especially for rural farmers. Nevertheless, adaptation and mitigation measures appropriate for agriculture are being promoted in developing countries of Asia which may be applicable in the Philippines. However, the role and support of national and local governments is critical in implementing these measures.

This paper describes the impacts of climate change in the agricultural sector in the Philippines. It also describes how these impacts influence the food security in rural areas, and discusses the socioeconomic implications. Finally, it suggests potential adaptation and mitigation measures in the agricultural sector and policy recommendations to combat climatic changes in the Philippines.

Impact of Climate Change on Agriculture and Food Security

A tropical country, Philippines has a dry season (northeast monsoon) from November to April and a wet season (southwest monsoon) from May to October. The country has four types of climates: Type 1 has a pronounced wet period (May to November) and dry period (December to April); Type 2 is characterized by no clear dry season with a maximum rainfall period from November to January; Type 3 has no distinct wet and dry periods but a relatively dry period from November to April; and Type 4 has more or less evenly distributed rainfall throughout the year (Lansigan 2005). Rural farmers depend on this climate pattern or type in cropping and management systems including seeding, cultivation, and harvest in their respective region. In addition, the country is located within the typhoon belt and subject to a number of natural hazards including typhoons, cyclones, landslides (normally resulting from typhoons and cyclones), volcanic eruptions, destructive earthquakes and tsunami. These extreme climate events adversely affect the agricultural sector and eventually the economy of the country.

23 ENSO – El Niño Southern Oscillation is a disruption of the ocean-atmosphere system that occurred in tropical

Pacific with significant impacts on global weather such as redistribution of rains with extreme flooding and droughts (Neelin et al 1998). El Niño refers to the drought and La Niña to flooding.

Page 301: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

301

Similar to other developing countries, agriculture in the Philippines is highly dependent on water availability and quality, soil fertility and land use. Water and soil continuum cannot be isolated as both are critical variables in crop production. The quality and quantity of water is affected by rainfall. Extreme cases of rain may lead to flooding or drought. During flooding, soil aeration is difficult if not impossible. Flooding can also trigger soil salinization from coastal water overflowing to the agricultural land. In addition, too much water could cause substantive crop losses due to loss of nutrients and soil erosion. On the other hand, rain-fed agriculture will be severely distressed during drought periods, and thus irrigation system becomes crucial. Meanwhile, periods of drought are aggravated by intense solar radiation leading to increased temperature and evaporation.

Studies conducted by the Climatic Research Unit of the World Wildlife Fund (CRU-WWF) in 1999 revealed that the annual mean temperature in the Philippines and its surrounding coastal waters is rising. At the same time, the percent of rainfall is decreasing according to data from the early 1980s and late 1990s (see Figure 1). In the 20th century, a 6 percent decline in rainfall was observed (CRU-WWF 1999). Peng et al. (2004) reported that rice yield declined with higher night time temperatures that are a result of global warming. The study reported an increase of 0.35oC and 1.13oC in the minimum and maximum night temperature, respectively, from 1979-2003. Furthermore, rice yield from irrigated field was examined in relation to this weather data. The analysis showed a decline of 10 percent grain yield for every 1oC increase in minimum temperature during the growing season, whereas effects of maximum temperature on rice grain yield was insignificant (Peng et al. 2004). Although plants may respond to warmer night temperature by acclimation or increase in respiration, other mechanisms such as tillering, leaf-area expansion, stem elongation, grain filling and crop phonological development during day-night differential may also mitigate grain yield reduction, and thus needs further investigation (Peng et al. 2004).

In contrast, severe drought, or El Niño, occurred in 1982-1983, affecting thousands of agricultural areas including reservoirs throughout the country. El Niño recurred in 1997-1998 (Figure 2), when the Philippines lost P3.2 billion24 in corn production (Lansigan 2005).

The CRU-WWF (1999) study noted the current trend toward intensely rising sea levels of 20 cm and 40 cm in Manila and Legazpi, as compared to a small increase before the 1960s. The study explained that excessive land reclamation and possible subsidence may have contributed, but that warming ocean waters along the coasts are also a likely cause.

Declining crop production due to climatic variability has severely affected the economy in the Philippines, which eventually will affect food security, particularly in the rural areas. Agricultural loss can worsen as the frequencies and recurrence of El Niño (extended dry season) and La Niña (heavy rainfall) is projected to increase and if mitigation strategies are not in place in these areas.

Households headed by women are found to be more vulnerable to the impacts of climate change compared to men, as are migrants versus native inhabitants in rural areas (Pulhin et al. 2006). The vulnerability of women-headed household is due to limited physical ability and overwhelming family burdens like caring for sick children or extreme events like crop failure that force them to borrow money for family subsistence. Migrants have difficulty obtaining or less access to new agricultural land to cultivate when moving to watershed areas that are government-owned and thus legally prohibited for encroachment or cultivation (Pulhin et al. 2006).

24 US$ 1= P38 averaged in 1997-1998 (source: oanda,com)

Page 302: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

302

Adaptation and Mitigation Policy Strategies in Agriculture

Jose and Cruz (1999) gave a comprehensive discussion on adaptation measures for agriculture as a response to climate change. The authors divided adaptation strategies into supply and demand parameters. Supply adaptation approaches include comprehensive watershed management and water allocation systems and procedure. Watershed degradation occurs due to logging and shifting cultivation, resulting in loss of soil fertility, erosion and siltation of streams, rivers and lakes. The Philippine government programs on watershed management aim to prevent further degradation of watershed, and strict implementation of the rules of these programs must be taken into action for this effort to succeed. On the other hand, systematic allocation of water is a response to the increasing demand of water, especially during drought period.

For demand adaptation approaches, enhancement of irrigation efficiency, introduction of low water use crops and efficient farming systems, recycling (reuse) of water, improvement of monitoring and forecasting systems for floods and droughts, and use of water pricing policies and structures were discussed by Jose and Cruz (1999). Irrigation accounts for most of the Philippines’ water consumption. Thus, efficient use of irrigation water needs to be considered, especially during extreme weather events. Adaptations may include changing the cropping schedule to decrease demand for irrigation during dry seasons or adjusting the cropping calendar according to excessive water availability for planting and growing periods (Lansigan 2003), lining canals to reduce water losses, maximizing efficient water use during abundant periods by constructing reservoir-type projects or redesigning irrigation facilities to reuse return flows, and introducing other water-saving techniques (Lansigan 2003)

Another demand adaptation is introducing low water-use crops and efficient farming practices. Agricultural biotechnology plays an important role in breeding crops efficient in water use as well as drought- or flood-tolerant crops. Lansigan (2003) suggested early maturing varieties and varying planting density for rice as tools to combat extreme weather events. In addition, traditional farming practices such as drip irrigation, mulching, and other improved irrigation methods as well as use of windbreaks to diminish wind speed and evapotranspiration can improve use of the dwindling irrigation water (Baradas and Mina 1996 in Jose and Cruz 1999).

The Philippine government has been encouraging recycling and reuse of water effluent in agriculture and industries. Jose and Cruz (1999) suggested drainage water reuse to extend supply of water for irrigation. More importantly, improvement in the monitoring and forecast systems for floods and droughts are critical for farmers. Intense drought or flooding can influence crop growth and yield. Further development and enhancement of the current system will provide crucial information to the farmers in advance of potential extreme climate events, giving them adequate time to prepare. Lansigan (2003) described that seasonal climate forecasts at the regional and local level provide useful information to farmers, extension workers, traders and others in rice production systems at all levels. In addition, water use management is crucial during these events. Because of the increasing scarcity of water, pricing policies and structures must be put in place. Traditionally water is regarded as a global public good. However intensifying demand and decreasing supply necessitates the incorporation of water pricing policies not only in agriculture but to other sectors as well.

Crop insurance is another approach to assist farmers during serious weather incidents. In the Philippines, cost of land preparation and rice crop establishment are the only variables covered in crop insurance (Lansigan 2005). The vulnerability of rice-growing areas in different locations with diverse climate types and risks from climate variability is not considered.

Page 303: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

303

Furthermore, the high insurance premium makes this less attractive to rural farmers. A more reasonable insurance premium plus an appealing coverage scheme that includes expected crop yield or expected revenue based on harvestable crop should be considered (Lansigan 2005).

Policy Options

The Philippines has initiated mainstreaming climate change initiatives in the country’s development plans. Donor agencies have recognized the importance of this and ensure that grants given include addressing climate change. Because of the country’s long experience with serious weather incidents and variability, some coping mechanisms are already in place. For example, upland farmers in Pantabangan watershed in the north have improvised and learned adaptation mechanisms as response to climate change (Lasco and Pulhin 2006).

A number of Philippine policies on natural resource management are in place, including Presidential Decrees, Republic Act, Executive and Administrative Orders (AO) and Senate Bills. The AO No. 171, passed in February 2007, created the Philippine Presidential Task Force on Climate Change. It is tasked to address and mitigate the detrimental impacts of climate change by leading the development of adaptations to these impacts. It will achieve this primarily by enhancing the capacity of institutions to address climate change (Lasco et al. 2008).

Lasco et al. (2008) discussed some policy recommendations to address climate change in the Philippines. This includes an aggressive and systematic approach to developing an information, education and communication campaign. Such a campaign could disseminate appropriate and correct information to improve understanding of the phenomenon and associated risks, and also could raise awareness of potential adaptation and mitigation strategies. Second, a participatory and multisectoral stakeholder approach for a concerted effort to combat climate change is necessary. All stakeholders – national, regional, local, private sector, NGOs and academic – must establish and enact partnership arrangements and effective operational links. Third, technology improvement on climate change as well as policy impact assessment of technologies or programs addressing climatic variability in the Philippines must be prepared. Finally, strong and honest political will and commitment supplemented with financial support for climate change activities is needed.

References

Asian Development Bank (ADB) 2008. Key Indicators for Asia and the Pacific 2008. 39th Edition. Special Chapter: Comparing Poverty Across Countries: The Role of Purchasing Power Parities. ADB, Mandaluyong, Philippines.

Baradas, M.W. and J.G. Mina. 1996. Water management 2000: Not by irrigation alone. Presented at Workshop on Food Security, Los Baños, Laguna, Philippines.

Jose, A. and N.A. Cruz. 1999. Climate change impacts and responses in the Philippines: water resources. Climate Research 12: 77-84

Jose, A.M., R.V. Francisco and N.A. Cruz. 1993. A preliminary study on the impact of climate variability/change on water resources in the Philippines. PAGASA, Quezon City.

Lasco, R.D, R. Gerpacio, P.A.J. Sanchez and R.J.P. Delfino. 2008. Philippine environment and climate change: An assessment of policies and their impacts. SEAMEO-SEARCA Policy Brief Series 2008-2.

Page 304: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

304

Lasco, R. and F.B. Pulhin. 2006. Assessing the role of watershed areas in mitigating climate change in the Philippines: The case study of the La Mesa Watershed. Journal of Environmental Science and Management 9(1): 19-29.

Lansigan, F. 2005. Coping with climate variability and change in rice production systems in the Philippines, pp 542-545. In K. Toriyama, K.L. Heong and B. Hardy (eds). Rice is Life: Scientific Perspectives for the 21st Century: Proceedings of the World Rice Research Conference, Tsukuba, Japan, 5-7 November 2004.

Lansigan, F. 2003. Assessing the impacts of climate variability on crop production and developing coping strategies in rainfed agriculture, pp 21-36. In S. Yokoyama R.N. Concepcion (eds). Coping against El Nino for Stabilizing Rainfed Agriculture: Lessons from Asia and the Pacific. Proceedings of a Joint Workshop, Cebu, Philippines, September 17-19, 2002.

Neelin, J.D., D.S. Battisti, A.C. Hirst, F.-F. Jin, Y. Wakata, T. Yamagata and S.E. Zebiak. 1998. ENSO theory. Journal of Geophysical Research. 103: 14261-14290

Peng, S., J. Huang, J.E. Sheehy, R.C. Laza, R.M. Visperas, X. Zhong, G.S. Centeno, G.S. Khush and K.G. Cassman. 2004. Rice yields decline with higher night temperature from global warming. PNAS 101 (27): 9971-9975

Pulhin, J.M., R.J.J. Peras, R.V.O. Cruz, R.D. Lasco, F.B. Pulhi and M.A. Tapia. 2006. Vulnerability of communities to climate variability and extremes: Pantabangan-Carranglan Watershed in the Philippines. http://www.aiaccproject.org/working_papers/Working%20Papers/AIACC_WP44_Pulhin.pdf

Rosegrant, M., C. Ringler, S. Msangi, T. Zhu, T. Sulser, R. Valmonte-Santos, and S. Wood. 2007. Agriculture and Food Security in Asia: The Role of Agricultural Research and Knowledge in a Changing Environment. J Semi-Arid Tropical Agricultural Research 4(1):1-35. http://www.icrisat.org/journal/specialproject/sp6.pdf

World Wildlife Fund-Climate Research Unit (WWF-CRU). 1999. Climate change scenarios for the Philippines. http://www.cru.uea.ac.uk/~mikeh/research/philippines.pdf

Page 305: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

305

Figure 1. Changes in (a) annual mean temperature (oC) and (b) annual precipitation (%) in the Philippines, 1900-2000. Changes are with respect to the average 1961-1990 climate values of 25.7oC and 2325 mm

Source: CRU-WWF. 1999

(b)

(a)

Page 306: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

306

Figure 2. The occurrence of El Niño during 1997-1998 in the Philippines

Source: PAGASA 2000, in Lansigan 2003

Page 307: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

307

Climate Change and Agricultural Development: Case Study in Tajikistan

Introduction

Tajikistan is the poorest country of Central Asia. About 42.8 percent of the population lives below US$2 a day and the GDP per capita is only US$355 (UNDP 2008). A recent publication mentions that most farmers of the country live on less than US¢2 a day (Reid, Simms, and Johnson 2007). According to estimates from the Food and Agriculture Organization of the United Nations (FAO) for 25 countries in Asia, Tajikistan had the highest proportion of undernourished people in the total population for the period 2003-2005, estimated at 34 percent (FAO 2008).

Tajikistan is a mountainous country in which glaciers in the central part of the country provide irrigation water during the dry and hot summer months. About 6 percent or around 8.4 thousand km2 of the country’s total area is occupied by glaciers (Tajik Met Service 2002). Glaciers are substantially important for livelihoods around the country as they are the primary source of clean drinking water, water for irrigation, and for generating electricity (IRIN 2007a).

Permanent pastures cover 23 percent of Tajikistan’s total area and arable land covers around 8 percent, of which about 70 percent is irrigated (FAO 2001). The vast majority of the population lives in rural areas. Despite the fact that the share of the agricultural sector in the GDP is only 25 percent, it employs about 70 percent of the population (The Republic of Tajikistan 2008). The rural population has to struggle to produce in scarce arable land—0.14 hectares per capita (ADB 1999)—that is susceptible to various erosion processes present in about 90 percent of the cropland area (Tajik Met Service 2002).

With such a mountainous territory, many disasters such as avalanches often happen. In fact, avalanches and other natural disasters make Tajikistan one of the most disaster-prone countries in the world. Every year about 50,000 landslides, 5,000 tremors and earthquakes, and hundreds of avalanches and debris flows happen in the country (IRIN 2007b) adversely impacting agriculture. In the period of 1991-2000, hydrometeorological disasters caused severe agricultural losses. Main losses related to disasters were caused by: high air temperatures, hot winds and low air temperatures (9-13 percent); showers, flash floods and mudflows (10 to 15 percent); and hurricanes and sandstorms (5 to 7 percent). In the same period, one third of overall agricultural losses were caused by extreme weather events. In years with droughts, lack of precipitation and small snow stocks, the harvest of cereals was 10 to 30 percent lower than in normal years (Tajik Met Service 2002).

Just like Uzbekistan, Tajikistan is one of the most corrupt countries in the world. Transparency International ranked Tajikistan 151 out of 180 countries in terms of corruption (the higher the rank the higher the corruption level). Corruption at all levels of the public sectors has been blocking efforts of donors such as the African Development Bank (ADB) (Marat 2006). Therefore, corruption and political instability are likely to increase vulnerability to climate change and threaten the effectiveness of adaptation strategies.

Impact of Climate Change on Agriculture and Food Security

According to data from the period 1940-2005, there was a warming trend in Tajikistan during the cold season (November-December) in the range of 1-3oC. Decreasing temperatures were observed in high altitudes in February, March, May, June, and October. A trend to cold spells was also observed during the spring in some sub-mountainous and mountainous areas. By 2030,

Page 308: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

308

an increase in mean annual temperature in the range of 0.2-0.4oC is expected for most regions of Tajikistan, with the highest increase expected for the winter (The Republic of Tajikistan 2008).

Due to the complexity of the mountain landscape there is medium and low confidence in precipitation scenarios (UNEP/GRID-Arendal 2002). Observed trends of precipitation throughout the country are not uniform. In 1961-1990, a reduction in the amount of annual precipitation in the range of 1-20 percent was observed in the mountains of Central Tajikistan, the valleys of southwest and northern Tajikistan, the foothills of the Turkestan range, and the Eastern Pamir. Increases in precipitation in the range of 12-18 percent were observed in: Karategin-Darvaz and Western Pamir, from the altitude of 1,500 meters and higher; and Western Pamir (UNEP/GRID-Arendal 2002).

Glaciers are already being adversely impacted by climate change. In the 20th century, glaciers in Tajikistan lost more than 20 km3 of ice. Small glaciers of up to 1 km2 (small glaciers are 80 percent of all glaciers corresponding to 15 percent of total ice covers) melted intensively (UNEP/GRID-Arendal 2002). Following this trend, thousands of small glaciers are very likely to disappear from Tajikistan with global warming. In the beginning, glacier melting might increase water availability, but in the mid and long-term, substantially decreases in water flow in many rivers are expected (UNEP/GRID-Arendal 2002).

Past observations show that climate can directly affect pasture productivity. For each temperature rise of 2-4oC in February and March, it was observed a decrease in winter-spring pastures productivity of 20 percent. In the mid-mountain areas, temperature increases of 1.5 to 2oC were also associated with a decrease in pasture productivity by 20 percent. On the other hand, in high mountain pastures, pasture productivity increased with temperature increases of 1.5oC to 3oC (Tajik Met Service 2002).

Climate change will also have different impacts on crop yields according to the region. For cotton yields, increases are expected in the Vakhsh and Gissar regions and decreases in the Kulyab and Sogd regions, according to a HadCM2 model (Tajik Met Service 2002). Cereal productivity is expected to increase in the rain-fed lands of Gissar, Karategin, and Kyzylsu regions and decrease in the Northern Turkestan and Western Pamir regions (Tajik Met Service 2002). However, as mentioned before, expected increases in extreme events will increase to negative impacts in all regions.

Coping/Adaptations Strategies

Farmers in Tajikistan already perceive changes in the climate. During interviews with non-profit organizations, the Tajik people have mentioned that they have observed greater extremes of temperature, with hotter summers and colder winters and less predictable snow and rainfall. In an interview made by Oxfam researchers, a member of the Kulyab emergency committee reported that in recent years, there have been more floods, mudflows, and melting snow from the high mountains (Reid, Simms, and Johnson 2007). During interviews made in the Varzob District in 2007, local farmers said that they were not able to grow wheat during recent years as a result of a shortening in the growing season and of cold weather beginning before the wheat harvest could be collected. More frequent heavy rains have also been observed, which have destroyed crops (IRIN 2007a).

Climate extremes are felt more intensely because of poor infrastructure. In March 2007, a drought in the Vose District in Southern Tajikistan left the region without hydropower/electricity and therefore there was less clean water, as water pumps were not working. After the drought, excessive rains caused mudslides that destroyed the water supply system (Reid, Simms, and

Page 309: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

309

Johnson 2007). A water pipe built by Oxfam that used to feed spring water for 3,000 people was completely destroyed. After the pipe broke, people expressed their anger, as they were “drinking dirt in a land seen as the water basin of Central Asia (Walsh 2007).” Overall, 40 percent of the countries’ population lack access to safe water (improved water source) (UNDP 2008).

On three villages in the Varzob District, the international organization CARE worked with community-based organizations to implement pilot adaptation strategies to address the impact on food security of the shifting winter season. In order to address the problem of a shortening growing season, cold frames (small greenhouses) were used to start seedlings in the spring and to extend the growing season later in the year. The project prioritized female-headed households who were given cold frames plus training on use and maintenance. Food preservation techniques were also promoted to increase food availability in the winter. The project trained female members of a local organization to preserve tomatoes, onions, peppers, and other kitchen garden vegetables (Reid, Simms, and Johnson 2007). As a result, some households have been able to grow cold-hardy vegetables all year round. Some women have also been able to sell surplus products, increasing the family’s income. Furthermore, some households have constructed cold frames with their own resources (CARE 2007).

Simple adaptation measures like the one above that involves the participation of local communities can be effective. During interviews in different regions of the country, locals have claimed that they have insufficient information about climate change and possible adaptation measures. According to Reid et al. (2007), this might be the result of relatively small number of environmental NGOs operating in the country and lack of prioritization of the issue by the Government of Tajikistan and the media.

References

--------. 2007b. Tajikistan: Earthquake and avalanches hit the country. IRIN—The humanitarian news and analysis service of the UN Office for the Coordination of Humanitarian Affairs. Available at <http://www.irinnews.org/Report.aspx?ReportId=71165>. Accessed March 11, 2009.

--------. 2008. The State of food insecurity in the world. Rome: FAO.

ADB. 1999. Country assistance plan (2000-2002). Tajikistan. Manila, Philippines: Asian Development Bank.

CARE. 2007. Care in Tajikistan: Women as agents of adaptation. Available at <http://www.nationalteachin.org/PDFs/CARE/Tajikistan%20Case%20Study.pdf>. Accessed March 10, 2009.

FAO. 2001. Seed policy and programmes for the Central and Eastern European countries, Commonwealth of Independent States and other countries in transition. FAO Plant Production and Protection Paper No. 168. Rome: FAO.

IRIN. 2007a. Tajikistan: Climate change threatens livelihoods of mountain villagers. Reuters AlertNet. Available at http://www.alertnet.org/thenews/newsdesk/IRIN/12ba71892268719599a1659079095fc8.htm Accessed March 10, 2009.

Marat, E. 2006. Tajikistan: Corruption mounts as presidential elections near. Eurasia Daily Monitor Volume: 3 Issue: 156. Available at http://nie.wikispaces.com/Tajikistan. Accessed March 11, 2009.

Page 310: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

310

Reid, H., A. Simms, and V. Johnson. 2007. Up in smoke? Asia and the Pacific. The threat from climate change to human development and the environment. London: NEF.

Tajik Met Service. 2002. Tajikistan 2002: Vital maps and graphics on climate change. UNEP/GRID-Arendal: Tajik Met Service. Available at http://enrin.grida.no/htmls/tadjik/vitalgraphics/eng/index.htm.

Tajikistan, State Agency of Hydrometereology. 2008. The Second national communication of the Republic of Tajikistan under the United Nations Framework Convention on Climate Change. Dushanbe: State Agency of Hydrometereology.

UNDP. 2008. Human development report 2007/2008—Country fact sheets—Tajikistan. United Nations Development Programme. Available at http://hdr.undp.org/en/statistics/. Accessed March 7, 2009.

UNEP/GRID-Arendal. 2002. Tajikistan 2002. State of the environment report. Arendal, Norway: GRID-Arendal.

Walsh, N. P. 2007. Melting glaciers hit Tajik lives. Channel Available at <http://www.channel4.com/news/articles/society/environment/melting%20glaciers%20hit%20tajik%20lives/548167> . Accessed March 10, 2009.

Page 311: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

311

Climate Change and Agricultural Development: Case Study in Thailand

Introduction

Thailand is located in Southeast Asia and has borders with Cambodia, Lao PDR, Myanmar, and Malaysia. It has a total area of 514,000 km2 comprised of a land area of 511,770 km2 and water area of 2,230 km2 (World Factbook 2009). The country’s current climate conditions are described as rainy, warm and cloudy due to the southwest monsoon from May to September; and dry and cool due to the northeast monsoon from November to March (World Factbook 2009). The main agricultural commodities are rice, cassava (tapioca), rubber, corn, sugarcane, coconuts, and soybeans.

In 2008, the agricultural sector contributed with 11.4 percent to the GDP (about US$ 553.4 billion at purchasing power parity in 2008) while the industry and service sectors contributed 44 percent (World Factbook 2009). Despite its relatively low contribution to the GDP, agriculture employs over 50 percent of the economically active population, so much of the population depends on agriculture for livelihood and as a result is highly sensitive to climate change. Furthermore, approximately 70,000 households depend on marine fish capture and aquaculture (World Bank 2006). Even in the absence of climate change, competition from commercial fisherman and declining fish stocks threaten the sustainability of this source of employment.

A recent study measures vulnerability to climate change in Southeast Asia countries. The Bangkok region of Thailand has a high overall vulnerability to climate change (Yusuf and Francisco 2009). Specifically, this region is vulnerable to sea level rise and floods, while the Southern region of Thailand is considered a drought “hot spot” (Yusuf and Francisco 2009).

In contrast, in terms of socio-economic indicators, Thailand seems to be more resilient to the impacts of climate change than some of its neighbors. A high GDP growth (5.5 percent), significant agricultural production growth (4.4 percent), and low unemployment (1.5 percent) (World Bank 2006) indicate that Thailand might be better able to adapt to climate change than nearby countries with poorer performance in socio-economic indicators such as Cambodia, Myanmar and Lao PDR. In addition, literacy rates are high (96 percent) and the proportion of the population that is undernourished has declined from 29 percent in 1990 to 17 percent in 2003. However, despite overall growth and improvement in human development indicators, pockets of poverty continue to exist in Thailand - about 10 percent of the Thai people live below the international poverty line (ADB 2008).

The remainder of this brief paper describes the impacts of climate change in the agricultural sector in Thailand. It will likewise describe how these impacts influence food security as well as the socioeconomic implications of climate change for rural areas. The final section illustrates potential adaptation and mitigation measures and policy recommendations to combat climatic change and its detrimental effects to agriculture.

Impact of climate change on agriculture and food security

Assessments vary on the impact of climate change on agriculture. A study shows that rice yields in sites located in the Ubonratchathani Province in the lower Mekong River basin in Thailand are likely to benefit from a warmer climate. Specifically, rice yields could increase between 3.07 and 5.96 percent, depending on the assumed atmospheric concentration of CO2 (Snidvongs 2006) The study did not, however, assess the impacts of more frequent and intense floods (Snidvongs

Page 312: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

312

2006). The number of Thai people exposed to flash flooding was nearly 3 million in 2005 (Sanker, Nakano, and Shiomi 2007), so there is need to expand research to include this type of disaster. Furthermore, higher precipitation levels are projected to increase the water level in the Mekong River’s tributaries in Lao PDR and Thailand, increasing the risk of flooding (Snidvongs 2006).

In contrast to the predictions on positive yield effects for rice within one region, other studies project that agricultural production capacity will decrease under climate change. Cline (2007) found that agricultural production capacity in Thailand would decrease between 15 and 26 percent in 2080, depending on the effects of carbon fertilization. Mendelsohn (2005) predicts similar negative trends in agricultural GDP; depending on the estimation technique, Thailand could experience losses between US$0.3 and 19.6 million per year in 2100.

Sea level rise is an ongoing concern in Thailand that will be exacerbated under a changing climate. Currently, 11 percent of the coastline area along the Gulf of Thailand is eroding at a rate of more than 5 meters a year (World Bank 2006). It is estimated that, at this rate, sea level rise would affect 12 percent of the Thai population, lower GDP by 23 percent, erode 4 percent of agricultural land, and 36 percent of wetlands (World Bank 2006). These impacts on GDP will be felt by farmers and fisher folk who depend on coastal land and waters for a source of livelihoods.

Adaptation and mitigation policy strategies in agriculture

A household survey conducted by Jarungrattanapong and Manasboonphempool (2008) gives evidence of the effectiveness of household adaptation strategies in the Bang Khun Thian district that depends on shrimp farming, and has been directly affected by coastal erosion. The households in this region adopt a number of strategies, including construction of flood protection infrastructure to protect current households, constructing new households or moving old ones further inland, and/or accommodating new sea levels by renovating existing structures (Jarungrattanapong and Manasboonphempool 2008). The researchers estimated that the average cost of adaptation measures was high: approximately 107,587 baht, or US$3,130 per household, an amount which accounts for 23 percent of the annual household income. Furthermore, the researchers calculated that the rise in sea level inundates on average 8 percent of aquaculture area for the households surveyed. These results demonstrate that despite the autonomous implementation of strategies to deal with the effects of rising sea level, livelihoods were still negatively impacted.

In another example in the Yasothorn province, one of Thailands’ ten poorest provinces, prolonged drought and unpredictable flooding were destroying rice crops. Families in the region learned about an adaptation program initiated by a local non-profit organization called Earth Net Foundation and Oxfam. The program consists of promoting organic jasmine rice farming which depend less on off-farm inputs and requires less energy and more environmentally sound, compared to conventional chemical-based farming. The practice was successful; soil fertility increased and families reported higher rice yields and much higher profits as they do not use chemicals. Furthermore, organic rice showed to be much more resistant to drought and water scarcity than chemically-grown rice crops. Local families mentioned that their livelihood has significantly improved after the program was implemented. As a result, they are better adapted to cope with more frequent and intense droughts (Oxfam 2009). Whether this program can be upscale is an important issue.

Page 313: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

313

Much can be done to help rural communities to better cope with climate change. According to the Office of Environmental Policy and Planning (2000), the agricultural sector has continuously adapted to different challenges and become more diverse. It has also become more environmentally friendly – soil conservation, reduction in the application of chemical pesticides and fertilizers, and chemical-free or organic agriculture have been promoted. Precautionary measures that could be adopted to improve the adaptive capacity of the agricultural sector are the following:

Conservation and improvement of local drought resistant varieties

Improvement of cropping practices to minimize water use

Application of risk averse cropping systems

Analysis of potential crop substitution in different regions

Promotion of crop diversification program (Office of Environmental Policy and Planning 2000).

Crop improvement strategies are essential to adapt Thai agriculture to floods and droughts likely to be more intense due to climate change. Currently, there is a dynamic program that aims to improve drought tolerance in rainfed lowland rice. About 76 percent of the total rice growing area of the country is comprised of rainfed agriculture, mostly in the Northeast and North of Thailand. Recent developments have been made to improve the breeding program which include: improvement of the selection process, identification of target domains, development of drought screening facilities, integration of marker assisted breeding and introduction of farmer participatory plant breeding into conventional breeding. The farmer participatory approach is to ensure that farmers will accept new cultivars (Jongdee et al. 2004).

The harmful impacts of climate change on agricultural production have serious repercussions to the livelihoods of small farmers and fisher folk, who are the most vulnerable to climate extreme events. Adger (1999) defined vulnerability as the exposure of individuals or groups to livelihood stress due to impacts of environmental changes. Vulnerability of individuals or groups to environmental stresses is influenced by institutional and economic factors. Farmers are more vulnerable to changing climatic conditions because of their limited livelihood options, resources and to some extent skills in securing other job. Policy options to combat climate change, improve agricultural production and improve the socioeconomic conditions of very poor and vulnerable groups, especially women, need to be implemented.

References

Adger, W. N., and M. Kelly. 1999. Social vulnerability to climate change and the architecture of entitlements. Mitigation and Adaptation Strategies for Global Change 4: 253-266

Asian Development Bank (ADB) 2008. Asian Development Bank and Thailand Fact Sheet. http://www.adb.org/Documents/Fact_Sheets/THA.pdf

Chinvanno, S. 2003. Information for sustainable development in light of climate change in Mekong River Basin. Southeast Asia START Regional Center, Bangkok, Thailand.

Chinvanno, S., S. Souvannalath, B. Lersupavithnapa, V. Kerdsuk, and N. T. H. Thuan. 2006. Climate risks and rice farming in the lower Mekong River countries. AIACC Working

Page 314: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

314

Papers No. 40. Washington, DC: Assessments of Impacts and Adaptations to Climate Change (AIACC).

Cline, W. 2007. Country-level agricultural impact estimates. In Global warming and agriculture: Impacts estimates by country. Washington, DC: Peterson Institute.

FAO. 2008. The state of food insecurity in the world. Rome.

Heffer, P. 2009. Assessment of Fertilizer Use by Crop at the Global Level. 2006/07 - 2007/08. International Fertilizer Industry Association. Paris.

Jarungrattanapong, R. and A. Manasboonphempool. 2008. Adaptation Strategies for Coastal Erosion/Flooding: A Case Study of Two Communities in Bang Khun Thian District, Bangkok. TDRI Quarterly Review (March): 11.

Jongdee, B., G. Pantuwan, S. Fukai, and K. Fischer. 2004. Improving drought tolerance in rainfed lowland rice: an example from Thailand. 4th International Crop Science Congrees. Brisbane, Australia

Mendelsohn, R. 2005. Climate change impacts on Southeast Asian agriculture. Yale University.

Office of Environmental Policy and Planning. 2000. Thailand's Initial National Communication. Bangkok: Ministry of Science, Technology and Environment

Oxfam. 2009. Thai farming community adapts to climate change as world leaders shun solutions. OXFAMhttp://www.oxfamblogs.org/eastasia/?p=505 (accessed April 27, 2009)

Sanker, S., H. Nakano, and Y. Shiomi. 2007. Natural disasters data book—2006: An analytical overview . Kobe, Japan: Asian Disaster Reduction Center.

Snidvongs, A. 2006. Vulnerability to climate change related water resource changes and extreme hydrological events in Southeast Asia. A final report submitted to Assessments of Impacts and Adaptations to Climate Change (AIACC), Project No. AS 07. Washington, DC: International START Secretariat.

World Bank. 2006. Thailand Environment Monitor 2006. Bangkok.

World Fact Book. 2009 https://www.cia.gov/library/publications/the-world-factbook/geos/th.html, accessed on April 15, 2009.

Yusuf, A. A., and H. Francisco. 2009. Climate change vulnerability mapping for Southeast Asia. South Bridge Court, Singapore: Economy and Environment Program for Southeast Asia (EEPSEA).

Page 315: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

315

Climate Change and Agricultural Development: Case Study in Uzbekistan

Introduction

Uzbekistan has a total area of 447,400 km2 divided in 12 provinces and the autonomous republic of Karakalpakstan. The country has three agro-ecological zones: the desert, steppe, and semi-arid region (60 percent of the country); the fertile valleys that skirt the Amu Darya and Syr Darya rivers, including the Fergana valley; and the mountainous areas in the east (Hakimov et al. 2007).

Socioeconomic indicators of the country indicate a high sensitivity of the agricultural sector and rural population to climate change and variability25. Around 70 percent of the Uzbekistan’s population is estimated to live in rural areas (Abdurakhmanovi and Marnie 2006). In 2007, 24.4 percent of the GDP came from the agricultural sector (World Bank 2008). Furthermore, the country is highly dependent on irrigation. Ninety-three (93) percent of the water used in the country is for agriculture and over 80 percent of the arable land is irrigated. As a result, in 2003, Uzbekistan was the country in the Asia Pacific region with the largest proportion of irrigated agricultural land over total arable land (FAO 2007). A high dependence on irrigation for food security makes Uzbekistan a country extremely vulnerable to variations on water resources availability.

Water mismanagement and land degradation are likely to increase the sensitivity of the country to climate change. Changes in water use in Uzbekistan in the last decades have created downstream summer water shortages and winter water excesses (Hakimov et al. 2007). Great quantities of water are lost through evaporation and filtration as a result of inadequate conditions of irrigation networks and poor water resources management. Losses in the irrigation network are estimated to be around 40 percent (Republic of Uzbekistan 2008). Furthermore, about 50 percent of Uzbekistan’s agricultural land area is prone to degradation. Natural pastures, biological diversity, and ecosystems in the Amu-Darya delta are all being significantly reduced. Salinity and soil erosion caused by widespread inappropriate drainage practices associated with poor irrigation threatens irrigated crop production (Hakimov et al. 2007). In the Republic of Karakalpakstan, in most parts of the territory it is possible to see salt on the ground left over from evaporation (Reid, Simms, and Johnson 2007). In general, it has been observed that the productivity of irrigated cultivated land has been progressively declining in Uzbekistan (Khusanov and Kosimov 2007). In Karakalpakstan, productivity of basic agricultural crops has decreased by 20-30 percent since the 1980s (Republic of Uzbekistan 2008).

The capacity of the country to adapt to climate change might also be compromised by growing food insecurity. Uzbekistan is considered a medium human development country with a human development index (HDI) of 0.70226, that ranked the country to 113 out of 117 (HDR 2009). Despite a medium HDI, undernourishment seems to be increasing in the country. Other countries in Central Asia such as Tajikistan and Armenia have a higher proportion of the undernourished in total population. However, according to data from the Food and Agriculture Organization of the United Nations (FAO), from 1990-1992 to 2003-2005, Uzbekistan has had

25 More information about the three elements of vulnerability (exposure, sensitivity and adaptive capacity) can be

found in chapter 3. 26 Human development index (HDI) gives a composite measure of three dimensions of human development: a)

living a long and healthy life (measured by life expectancy); b) education (adult literacy and enrolment at the primary, secondary and tertiary level) and c) decent standard of living (purchasing power parity, PPP, income). The index is provides a broadened prism for viewing human progress and the complex relationship between income and well-being (HDR 2009).

Page 316: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

316

one of the largest proportional increases of the undernourished in the total population, from 5 to 14 percent. This increasing trend goes in the opposite direction of the general Asian trend of a reduction in undernourishment rates for the same period, from 20 to 16 percent (FAO 2008). In 2001, around 27.5 percent of the population was living below the poverty line and 9 percent in extreme poverty (Abdurakhmanovi and Marnie 2006).

Another possible obstacle for the successful adaptation of the agricultural sector to climate change is the fact that many poor farmers barely make a living out of the land owned by the government. In the 1990s, small plots of lands (averaged size of 0.12 hectares) owned by the state were allocated to the rural and peri-urban population. An estimated 82 percent of Uzbekistan’s households benefited from these plots, which accounted for 80 to 90 percent of meat and milk production in the country and half the agricultural GDP. However, the plots were in many cases enough only for subsistence purposes (Abdurakhmanovi and Marnie 2006; UNDP 2000). Therefore, high poverty rates, despite good growth rates and low unemployment rates, can in part be explained by a large number of underemployed engaged in low-productivity and low-income jobs particularly in the agricultural sector (Abdurakhmanovi and Marnie 2006).

On the top of social problems, Uzbekistan is considered one of the most corrupt countries in the world. In the 2008 corruption report published by the civil society organization Transparency International, Uzbekistan ranked 166 out of 180 countries (the higher the rank the higher the corruption level). Climate change might contribute to an increase in corruption. The adverse impacts of climate change on the water sector, for instance, might lead to more corruption, as new water infrastructures such as the ones related to flood control might be needed giving opportunities for corruption through contracts or may even save the funds for personal use. Furthermore, powerful segments of the society will have incentives to have water control, supplying water through briberies and political lobbying (Transparency International 2008).

Therefore, high undernourished rates, high dependency on irrigated agriculture, land degradation, and high corruption rates make the country more sensitive and less capable to adapt to climate change, increasing its overall vulnerability.

Impacts of climate change on agriculture

Recent estimations project that annual mean temperature will increase in Uzbekistan from 12.51oC (1961-90) to 14.55oC (2070-99). On the other hand, mean precipitation is projected to decrease (IFPRI estimate). Climate change might also lead to an increase in the incidence of extreme events intensifying the impact of long droughts during the summer and frosts in late spring and early fall that frequently happen in Uzbekistan causing losses in crop production (Hakimov et al. 2007).

According to estimates made by the Government of Uzbekistan for the Second National Communication report, by 2050, water flow is expected to decrease by 2-5 percent in the Syrdarya River Basin and by 10-15 percent in the Amudarya River Basin under the A227 scenario (Republic of Uzbekistan 2008). In extremely warm and dry years, with high water scarcity, vegetation flow in those two basins might decrease by 25-50 percent. According to this report, crop losses in the future will be mainly caused by changes in water security of irrigated farming. Furthermore, despite the fact that some scenarios project that Uzbekistan will not have a water availability problem in the future, excessive use of water flow from the major rivers as a

27 A2 scenario is described by lower trade flows, relatively slow capital stock turnover, and slower technological

change (IPCC 2001).

Page 317: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

317

result of projected increases in the number of people living in the Amu-Darya and Syr-Darya river basins might increase water stress in the region (Hakimov et al. 2007).

Cotton is the main strategic agricultural crop for the Government of Uzbekistan. Uzbekistan is the world’s second largest cotton exporter; in 2006, the production of raw cotton was over 3.6 million tons. As a result of evaporation, cotton crops production potential might decrease by 4 percent in 2030 and 10 percent by 2050. In terms of food security, wheat is the most consumed food in the country; the per capita calorie intake is 1207 calories (FAO 2007a). For winter wheat, yields might decrease by 2 percent in 2030 and 4 percent in 2050. In extreme years, losses for the majority of crops could go up to 14 percent (Republic of Uzbekistan 2008).

Emissions

Uzbekistan accounts for 0.5 percent of the world’s total emissions, which is slightly higher than its share of the world’s population (0.4 percent) (UNDP 2008). In 2000, the agricultural sector was responsible for 8 percent of total emissions of the country. In the period of 1990-2005, methane emissions increased by 33.1 percent, which is related to an increase of livestock and sheep herds. On the other hand, nitrogen emissions declined by 26.4 percent, due to a decreased utilization of the mineral fertilizers (Republic of Uzbekistan 2008).

Adaptation Strategies

In general terms, the Government of Uzbekistan perceives the following measures as main adaptation strategies that should be done in the agricultural and water sectors (Republic of Uzbekistan 2008):

Improve land and water resources management at national and transboundary level

Save water and rationally use water in irrigated land

Improve water resources monitoring system

Save water in industrial and household water consumption

Increase plant growing productivity (which includes the introduction of drought-resistant species)

Increase animal husbandry productivity

However, the Uzbek Government acknowledges that those strategies face several barriers: insufficient regional coordination; insufficient financial resources and investment; and insufficient applied research, development, and technical expertise. Management of transboundary water resources is particular relevant to Uzbekistan, as about 90 percent of the region’s total annual river run-off is formed beyond its boundaries (Hakimov et al. 2007).

Adaptation measures aimed at improving the monitoring of climate extreme events and lakes prone to bursting are also essential, as such events are expected to increase with climate change. However, successful hydrometeorological monitoring is also compromised by insufficient regional coordination, lack of applied research and developments, and shortage of specialists (Republic of Uzbekistan 2008).

In order to better cope with climate variability, some water management strategies have taken place in the transboundary Amudarya basin (about 8.5 percent of Amu Darya water is formed in Uzbekistan, however, the country uses a much higher quantity of its water) (Krysanova et al. 2008). The main strategies are listed below:

Page 318: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

318

Two single-year storage reservoirs, small-scale reservoirs in the wetlands of the river delta, and the use of former bays of the Aral;

Sea for water storage, and water-allocation planning measures, e.g., reduction of allocation quotas to all regions by fixed percentages when a low-water year is forecast;

Water-saving technologies related to irrigation to decrease water demand in agriculture and increase efficiency; and

Establishment of a transboundary flood emergency committee to ensure quick responses to flood risks; and campaigns to raise awareness among water users about water conservation and flood protection.

However, according to Krysanova et al. 2008, some recent extreme events have revealed that those measures might be insufficient to cope with more intense events in the future. Vulnerability of the basin is exacerbated by lack of maintenance and investments and overuse of water resources. Therefore, measures aimed to improve water efficiency use and reduce waste, including financial and legal incentives as well as more transboundary cooperation are essential for the development of successful coping strategies (Krysanova et al. 2008).

Therefore, much has yet to be done to increase the capacity of the Uzbek people to better cope with climate change. It is also important to understand how the Uzbek rural communities are coping with climate change and variability. Current coping mechanisms and adaptation strategies used in the country should be publicized so good strategies can be replicated and mistakes can be learned from unsuccessful ones. Unfortunately, there is not enough research on such matters.

As a way to increase the population’s awareness to climate change, the Center of Hydrometeorological Service at the Cabinet of Ministers of the Republic of Uzbekistan (Uzhydromet), (responsible for the implementation of the United Nations Framework Convention on Climate Change in Uzbekistan) created a site specialized in climate change (www.climate.uz) called Climate School. Overall, however, much more research needs to be done to access current and future impacts of climate change on food security in Uzbekistan.

References

--------. 2007. Selected indicators of food and agricultural development in the Asia-Pacific Region 1996-2006. RAP PUBLICATION 2007/15Rome: FAO.

--------. 2008. The state of food Insecurity in the World. Rome: FAO

Abdurakhmanovi, U., and S. Marnie. 2006. Poverty and inequality in Uzbekistan. Development and Transition Newsleter.UNDP, London School of Economics, London.

FAO (Food and Agriculture Organization of the United Nations). 2007a. FAO statistical yearbook - Countries profile - Uzbekistan. Rome: FAO.

Hakimov, A., A. Lines, P. Elmuratov, and R. Hakimov. 2007. Climate change and water resource alteration in Central Asia: the case of Uzbekistan. In Climate Change and Terrestrial Carbon Sequestration in Central Asia. Ed. R. Lal, M. Suleimenov, B. A. Steward, D. O. Hansen, and P. Doraiswamy Routledge USA.

IPCC 2000 Special Report on Emissions Scenarios. Nebojsa Nakicenovic and Rob Swart (Eds.). Cambridge University Press, UK. pp 570

Page 319: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

319

Khusanov, R., and M. Kosimov. 2007. Problems and management of the efficient use of soil-water resources in Central Asia with specific reference to Uzbekistan. In Climate Change and Terrestrial Carbon Sequestration in Central Asia. Ed. R. Lal, M. Suleimenov, B. A. Steward, D. O. Hansen, and P. Doraiswamy Routledge, USA.

Krysanova, V., H. Buiteveld, D. Haase, F. F. Hattermann, K. van Niekerk, K. Roest, P. Martinez-Santos, and M. Schluter. 2008. Practices and lessons learned in coping with climatic hazards at the river-basin scale: floods and drought. Ecology and Society 13 (2).

Reid, H., A. Simms, and V. Johnson. 2007. Up in smoke? Asia and the Pacific. The threat from climate change to human development and the environment. London: NEF.

Republic of Uzbekistan, Centre of Hydrometeorological Service. 2008. Second national communication of the Republic of Uzbekistan under the United Nations Framework Convention on Climate Change. Tashkent: Centre of Hydrometeorological Service.

Transparency International. 2008. Global corruption report 2008. Corruption in the water sector. Cambridge: Cambridge University Press.

UNDP. 2000. Uzbekistan country assessment. UNDP poverty report 2000. New York: UNDP.

UN Human Development report 2008. Human development report 2007/2008 - Country Fact Sheets - Uzbekistan. UNDP. Available at http://hdr.undp.org/en/statistics/ . Accessed March 7, 2009.

World Bank. 2008. Uzbekistan at a glance. The World Bank. Available at http://devdata.worldbank.org/aag/uzb_aag.pdf. Accessed March 10, 2009.

Page 320: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

320

Climate Change and Agricultural Development: Case Study in Viet nam

Introduction

Viet Nam is one of the developing countries that will be severely affected by climate changes. Oxfam International (2008) reported that Viet Nam is one the top 10 countries that will be affected by climatic variability. Viet Nam significantly reduced its poverty rate from 58 percent in 1993 to 18 percent in 2006 through strong economic growth, pro-poor agricultural development policies and solid government commitment (ADB 2008a). It is one of the few countries on track in achieving the first Millennium Development Goal (“Eradicate extreme poverty and hunger”) by 2015 (Oxfam International 2008).

Viet Nam has a total land area of 320,000 sq. km and coastline of 3,260 km (Chaudhry and Ruysschaert 2008). Its land boundaries are China in the northeast, Laos in the central-west and Cambodia in the south. There are two major rivers in the country, the Red River Delta and the Mekong River Delta, in the north and south, respectively. Because of fertile soil and dense population in the southern region, agriculture and industries are the main economic activities in the region (Chaudhry and Ruysschaert 2008).

Viet Nam has a tropical monsoon climate which varies from north to south due to the length of the country and diverse topography (Chaudhry and Ruysschaert 2008). Flooding is the most common natural disaster, particularly in the southern part of the country, the Mekong Delta. Other environmental issues that could aggravate the current climatic conditions range from logging and slash-and-burn agriculture that leads to deforestation and soil erosion; increased turbidity from river run-off resulting in pollution and threatening freshwater and marine life; groundwater contamination that limits potable water supply; and growing urbanization, industrialization and urban migration, degrading the environment in major cities like Hanoi and Ho Chi Minh.

In 2006, Viet Nam had an estimated GDP per capita of US$2,363 (ADB 2008b). Industries, services and agriculture sectors at 41.5 percent, 38.1 percent, and 20.4 percent respectively contributed to the 2006 GDP (ADB 2008b). Significant improvement in the economic condition of the rural poor was achieved through the strong efforts of the government (Chaudhry and Ruysschaert 2008). However the strengthening socioeconomic conditions in rural areas will be at risk with the onset of changes in climatic conditions that affect the agricultural sector in the rural areas.

The changing pattern of typhoons and other natural hazards have severely affected the three main regions (north, central and south) of Viet Nam. These disasters do not only destroy agricultural harvest but claim lives as well. More than 8,000 people were killed due to natural disasters like storms, flood, flash floods and landslides between 1991 and 2000 (Chaudhry and Ruysschaert 2008). Table 1 shows the disaster zone and hazards in Viet Nam while Figure 1 illustrates these in map form.

This paper describes the impacts of climate change within the agricultural sector in Viet Nam and explains how these impacts influence the food security and socioeconomic situation in rural areas. Finally, it illustrates potential adaptation and mitigation measures and policy recommendations to combat climatic changes with detrimental effect to agriculture in Viet Nam.

Impact of Climate Change on Agriculture and Food Security

Viet Nam’s annual mean temperature ranges from 18oC to 29oC, with cooler temperatures in the mountainous area in the northern region (Chaudhry and Ruysschaert 2008). An increase of

Page 321: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

321

0.1oC-0.3oC in the annual average temperature per decade from 1900 to 2000 was reported by Chaudhry and Ruysschaert in 2008. This elevation in temperature resulted in hotter summer months. Furthermore, a more pronounced increase of 1.4oC-1.5oC by 2050 and 2.5oC-2.8oC by 2100 is expected (Hoang and Tran 2006). Nguyen et al. (2005) estimated a fall of 3 percent to 6 percent in summer rice yield in 2070 compared to the 1960-1998 level. Projected declines for spring rice are even more severe, with a predicted drop of 17 percent in yield in 2070 in the north and at 8 percent decline in the south (Chaudhry and Ruysschaert 2008). Maize production will likewise suffer from the impacts of climate change. A 4 percent and 9 percent decrease in yields of spring maize was estimated for central and south Viet Nam, respectively (Nguyen et al. 2005). Interestingly, a 7 percent increase in the yield of spring maize was estimated in the north part of the country in 2070 (Nguyen et al. 2005), which would be a rare positive outcome of climate change.

Temperature variability alters the growing periods, crop calendars and distribution, and enhances incidence of pests and virus activities (Chaudhry and Ruysschaert 2008). Rising temperatures will result in declining agricultural production, thus affecting the socioeconomic condition of the farmers. Moreover, climate change also encourages deforestation. The Ministry of Natural Resources and Environment (MoNRE) pointed out that as a response to increasing minimum temperature, farmers will likely migrate to the hills for crop cultivation to ensure food availability and income (MoNRE 2003).

Livestock production will likewise be affected by temperature variability. In early 2008, a cold spell hit northern Viet Nam. The serious drop in temperature below 10oC and even as low as -2oC resulted in the deaths of more than 60,000 cattle (Chaudhry and Ruysschaert 2008). Additionally, at least 100,000 hectares of rice were destroyed, resulting in economic losses of around US$30 million (Chaudhry and Ruysschaert 2008).

The Ministry of Natural Resources and Environment has reported irregular rainfall patterns throughout the country (MoNRE, 2003). Precipitation varies from 1,400mm to 2,400mm on average but may be higher in other regions, which may cause flash floods, landslides and flooding (MoNRE 2003). The annual total rainfall is expected to increase by 2.5-4.8 percent in 2050 and 4.7-8.8 percent in 2100 compared to 1990 (Nguyen 2006). Hoang and Tran (2006) observed that most of the rain will fall in the north, where upland communities will face higher risk of flash floods and landslides due to typhoons. Wassmann et al. (2004) reported that excessive flooding in tidally inundated areas and longer flooding periods will adversely affect rice production including cropping seasons, namely, mua (main rainfed crop), dong xuan (winter-spring crop) and he thu (summer-autumn crop) in the delta.

Intense rain resulted in Ho Chi Minh City’s worst high tides ever in 2007. This led to the destruction of around 40 sections of dykes throughout the city (Oxfam International 2008). In contrast, the less rain in the southern plains aggravated crop production and even fisheries resources in the delta (Chaudhry and Ruysschaert 2008).

Aside from extreme flood and drought, sea-level rise is another problem that needs to be addressed, as this has significant impact in the delta. An increase of 3 cm to 5 cm in sea level was observed from the 1960s to the 1990s, although with regional variation (Oxfam International 2008; UNEP 1993). The intrusion of seawater leads to elevated levels of salt in the agricultural land which has detrimental effects on the crops. Furthermore, encroachment of water erodes coastal land. This has already occurrred in Cau Mau province, where more than 600 ha of land have been washed away due to the rise in sea level (Chaudhry and Ruysschaert 2008). Furthermore, intrusion of seawater will severely affect low-lying areas in the delta. It encroaches

Page 322: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

322

on freshwater availability for drinking and hinders extraction for irrigation and construction of canals in the deltas and dams in the upstream of Mekong River (MHC et al. 2006).

Although both men and women will be affected by natural disasters, women in the villages suffer more from climate change. Many women in rural Viet Nam cannot swim endangering their lives during flash flood. In addition, most women have fewer assets than men. Hence they have fewer options or alternative livelihoods when crops are destroyed by extreme weather. Further, expectations about their domestic roles and responsibilities may constrain their ability to seek other employment far from home (Oxfam International 2008).

Seawater inundation, storms, drought and flooding will have damaging impacts to crop production, particularly rice in the southern part of the country. Viet Nam ranks second in world rice exports, and climate change that undermines rice production will severely undermine the economy. This may lead to instability of farmers’ income and threatens food security, especially in the rural areas.

Adaptation and Mitigation Strategies in Agriculture

The harmful impacts of climate change to agricultural production have serious repercussions to the livelihood of the rural farmers, who are most vulnerable in extreme events. Adger (1999) defined vulnerability as the exposure of individuals or groups to livelihood stress due to impacts of environmental changes. Vulnerability of individuals or groups to environmental stresses are influenced by institutional and economic factors. Farmers are more vulnerable to changing climatic conditions because of their limited livelihood options, resources and to some extent skills in securing other work. Of all disasters resulting from climate change, the most catastrophic will result from rising sea levels, as Viet Nam was found to be at highest risk of the 84 coastal developing countries (Waibel 2008). High density populations and strong economic activities are concentrated in low-lying areas of the Mekong and Red River Deltas. A predicted 1-m rise in sea level will flood 20,000 km2 and 5,000 km2 of the Mekong and Red River Deltas, respectively, severely affecting population, agriculture and livelihood (Waibel 2008). Other extreme climate events such as flash floods, typhoons, drought and others will lead to similar impacts.

Communities and especially the farmers have experienced and are aware of these damaging impacts of climate change in crop production. In response, farmers have developed local-level coping strategies to ensure protection and production of crops. These strategies include changing seed varieties and crops, diversifying to non-farm techniques and seasonal migration (Chaudhry and Ruysschert 2008). At the community level, villagers worked together in protecting common property resources like fishponds while at the national level, the focus is on infrastructure investments, R&D and strengthening information systems (Suppakorn et al. 2005).

The government of Viet Nam has taken strong action in alleviating the impacts of climatic variability. Adaptation and mitigation strategies to lessen the negative impacts of climate change are embedded in development strategies of the country. Some of these strategies are discussed below.

Several adaptation strategies to combat climate change were proposed by the government of Viet Nam. In the coastal areas, 5,000 km of river dykes and 2,000 km of sea dykes were put in place as protection from typhoons and rising sea levels (Chaudhry and Ruysschaert 2008). Both the national and local governments are responsible in maintaining these dykes, and extensive labor was provided by the communities (Chaudhry and Ruysschaert 2008). Aside

Page 323: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

323

from protection against rising sea levels, collective action and trust among the national-local government and communities were the benefits gained from working together. Mangrove reforestation was also implemented to prevent coastal erosion and serves as wind breaker during typhoon season. Early warning systems for disasters were being improved to ensure on-time delivery of the information to the leaders and members of the community. Awareness on disaster risk management was similarly raised among community members. National departments, social and civic groups and other mass organizations were mobilized to distribute literature, organize meetings, visit households and inform/warn them of impending rising of flood waters and other potential calamities (Chaudhry and Ruysschaert 2008). Past devastating floods, where the majority of victims were poor children, was a lesson learned the hard way in Viet Nam. Hence the government and community worked and acted together to minimize the catastrophic impacts of these extreme natural disasters. One example of this collective action was the project on “community-based adaptation to climate change in Viet Nam” implemented in 2002 in four communes and eight villages in Quang Dien and Phu Vang Districts, located in Thuan Thien province along the north central coast of the country (Francisco 2008). A 1999 flood took hundreds of lives, as well as property and all means of livelihood. As part of the relief operations to assist the affected families in rebuilding their lives, a program on “capacity building for adaptation to climate change” was initiated. The main objective is to help build communities’ adaptive strategies to recurrent climatic catastrophes and minimize loss of lives and property should an event occur. The program was implemented in three phases: i) scenario building that identified and analyzed the hazards, vulnerability to climate change and existing and required adaptive capacity of respective villages; ii) planning that involved discussions among leaders of social groups or organizations on threats and potential impacts from climate change and adaptive measures to address these impacts; and iii) project implementation of sub-projects identified in the plan, which were made possible through in-kind and cash contributions from the community adaptation plans. Throughout the entire process, the government played a major role in providing technical expertise, guidance and ensuring timely implementation of the sub-projects.

Policy Options

The impacts of climate change have serious repercussions on socioeconomic development, particularly among the rural poor communities of Viet Nam. Chaudhry and Ruysschaert (2008) pointed out that despite the threats of climatic variability in the country, information exchange and awareness, especially at the community level, is poor. National or local climate change adaptation strategies are not yet in place, and capacity-strengthening of local and national agencies is urgently needed. Although action on adaptation strategies and policies still need to be strengthened, the government of Viet Nam has initiated coordination among line agencies to mainstream climate change in their development policies. The Ministry of Natural Resources and Environment is the national focal agency responsible for any activities related to climate change (Chaudhry and Ruysschaert 2008). The Ministry ensures that adaptation measures are incorporated in laws and national strategies. Some examples are the 2005 National Strategy for Environmental Protection that includes measures to reduce the impact of sea level rise in coastal zones; and the International Support Group on Natural Resources, based in MoNRE, which established a climate change adaptation working group. This working group promotes forum for dialogue and coordination for implementation of climate change adaptation measures (Chaudhry and Ruysschaert 2008). Other committees were instituted, such as the Central Committee for

Page 324: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

324

Flood and Storm Control (CCFSC) in 1995, chaired by the Ministry of Agriculture and Rural Development. Ministries like the Department of Floods and Storm Control and Dyke Management, Disaster Management Centre, Hydro-meteorological Service and Viet Nam Red Cross as well as the Natural Disaster Mitigation Partnership and others are the line agencies and members of CCFSC. Local CCFSC members at the provincial, district and commune levels are all active members of the CCFCS. These local level organizations are responsible for coordination of flood and storm measures, as well as organizing dyke protection and flood and storm preparedness and mitigation (EU/MWH 2006). The CCFCS are also responsible for providing early warning information, technical assistance, materials and equipment. Given these various initiatives, the government of Viet Nam is clearly aware of problems triggered by climate change and is taking necessary steps to mainstream appropriate responses into their development policies. Determining the most effective method to communicate and implement these policies will be a challenge not only in Viet Nam but in other developing countries as well. Technical expertise, financial support and willingness from all sectors and the communities to work together are necessary requirements to ensure agricultural productivity and protection of the people to mitigate the adverse impacts of climate change.

References

ADB 2008b. Key Indicators for Asia and the Pacific. 39th Edition. Special Chapter: Comparing Poverty Across Countries: The Role of Purchasing Power Parities. ADB, Mandaluyong, Philippines.

Adger, W.N. 1999. Social Vulnerability to Climate Change and Extremes in Coastal Viet Nam. World Development 27(2): 249-269

Asian Development Bank (ADB) 2008a. Asian Development Bank and Viet Nam Fact Sheet. http://www.adb.org/Documents/Fact_Sheets/VIE.pdf

Chaudhry, P. and G. Ruysschaert 2008. Climate change and human development in Viet Nam. Human Development Report 2007/2008. HDR 2007/46.

European Union (EU)/MWH 2006. Linking climate change adaptation and disaster risk management for sustainable poverty reduction. Viet Nam Country Study. Government Statistical Office. Statistical Yearbook of Viet Nam.

Francisco, H.A. 2008. Adaptation to climate change: Needs and opportunities for Southeast Asia. ASEAN Economic Bulletin 25 (1): 7-19

Hien Thuan 2006. Climate risks and rice farming in the lower Mekong River countries. AIACC Working Paper No. 40

Hoang, D.C. and V.L. Tran. 2006. Developing various climate change scenarios of 21 century for regions of Viet Nam, Scientific and Technical Hydro-Meteorological Journal No 541, January 2006.

MHC, Polish Academy of Sciences, and National Institute for Coastal and Marine Management (The Netherlands). 1996. Viet Nam Coastal Vulnerability Assessment: First Steps Towards Integrated Coastal Zone Management. Marine Hydrometeorology Center (MHC), Hanoi.

Page 325: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

325

Ministry of Natural Resources and Environment (MoNRE) 2003. Viet Nam Initial National Communication under the United Nations Framework Convention on Climate Change. MoNRE, Ha Noi, Viet Nam.

Nguyen, Le Tuong. 2006. Climate Change and Activities in Viet Nam.

Nguyen, M.C., V.H. Ninh, T.G. Ngo 2005. Study on Climate Change Impacts to Viet Nam Agriculture and Adaptation Measures. Technical paper prepared for the national programme on studying climate change impacts.

Oxfam 2008. Viet Nam: Climate change, adaptation and poor people. http://www.oxfam.org.uk/resources/policy/climate_change/downloads/Viet Nam_cc_adaptation_poverty.pdf

Suppakorn Chinvanno, Soulideth Souvannalath, Boontium Lersupavithnapa, Vichien Kerdsuk, and Nguyen Thi Hien Thuan. 2006. Climate risks and rice farming in the lower Mekong River countries. AIACC Working Paper No. 40

UNEP 1993. Viet Nam and Climate Change. Fact sheet produced by the Information Unit on Climate Change, United Nations Environment Programme, Geneva.

Waibel, M. 2008. Implications and challenges of climate change for Viet Nam. Pacific News 2(9): 26-27 (January/February 2008).

Wassmann, R., Nguyen Xuan Hien, Chu Thai Hoanh, and To Phuc Tuon. 2004. Sea Level Rise Affecting the Viet Namese Mekong Delta: Water Elevation in the Flood Season and Implications for Rice Production. Climatic Change 66: 89-107.

Page 326: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

326

Table 1. Climate related natural hazards by region in Viet Nam Region Disaster Zone Major Natural Hazards

North Northern uplands Flash floods, landslides, earthquakes

Red River Delta Monsoon river floods, typhoons, coastal storm surges

Center Central coast provinces Typhoons, storm surges, flash floods, drought, saline water intrusion

Central highlands Flash floods, landslides South Mekong River Delta River flooding, typhoons,

high tides and storm surges, salt water intrusion

Source: Chaudhry and Ruysschaert 2008

Page 327: Building Climate Resilience in the Agriculture Sectorotsuka/PPT2009/ADB Report - April 27 2009.pdfStrategies to build resilience including: Minimize reactive coping and promote proactive

327

Figure 1. Climate related natural hazards by region in Viet Nam (map form)

Source: Oxfam International 2008