Climate change in Africa

Anthropogenic climate change is already a reality in Africa, as it is elsewhere in the world. According to the Intergovernmental Panel on Climate Change, Africa is among the most vulnerable continents to climate change.[1][2] The vulnerability of Africa to climate change is driven by a range of factors that includes weak adaptive capacity, high dependence on ecosystem goods for livelihoods, and less developed agricultural production systems.[3] The risks of climate change on agricultural production, food security, water resources and ecosystem services will likely have increasingly severe consequences on lives and sustainable development prospects in Africa.[4] Managing this risk requires integration of mitigation and adaptation strategies in the management of ecosystem goods and services, and the agriculture production systems in Africa.[5]

Africa map of Köppen climate classification.

Over the coming decades, warming from climate change is expected across almost all the earth's surface, and global mean rainfall will increase.[6] Regional effects on rainfall in the tropics are expected to be much more spatially variable and the sign of change at any one location is often less certain, although changes are expected. Consistent with this, observed surface temperatures have generally increased over Africa since the late 19th century to the early 21st century by about  1 °C, but locally as much as 3 °C for minimum temperature in the Sahel at the end of the dry season.[7] Observed precipitation trends indicate spatial and temporal discrepancies as expected.[8][2] The observed changes in temperature and precipitation vary regionally.[9][8]

In terms of adaptation efforts, regional-level actors are making some progress. This includes the development and adoption of several regional climate change adaptation strategies[10] e.g. SADC Policy Paper Climate Change,[11] and the adaptation strategy for the water sector.[12] In addition, there has been other efforts to enhance climate change adaptation, such as the tripatite Programme on Climate Change Adaptation and Mitigation in Eastern and Southern Africa (COMESA-EAC-SADC).[13]

As a supranational organisation of 55 member states, the African Union has put forward 47 goals and corresponding actions in a 2014 draft report[14] to combat and mitigate climate change on the continent. The Secretary General of the United Nations has also declared a need for close cooperation with the African Union in order to tackle climate change, in accordance with the UN's sustainable development goals.

Impacts

Climate change will increasingly impact Africa due to many factors. These impacts are already being felt and will increase in magnitude if action is not taken to reduce global carbon emissions. The impacts include higher temperatures, drought, changing rainfall patterns and increased climate variability. These conditions have a bearing on energy production and consumption. The recent drought in many African countries, which has been linked to climate change, adversely affected both energy security and economic growth across the continent.

The possible effect of temperature rise caused by climate change on humans. The black stains, are the areas in which the mean annual temperature in the year 2020 is higher then 29°C, what is considered as too hot for normal life. The shaded areas, are the areas in which it will be higher than 29°C in the business as usual (RCP8.5) scenario by the year 2070[15] e.g., if the global average temperature will rise by 3.2°C, relatively to the pre - industrial baseline[16]

Agriculture and food security

Agriculture is inherently sensitive to climate conditions and is one of the most vulnerable sectors to the risks and impacts of global climate change.[17] Agriculture in most  African countries is mainly small-scale and rain-fed, making it particularly vulnerable to climate variability and change. Observed and projected disruptions in precipitation patterns due to climate change are likely to shorten growing seasons and affect crop yield in many parts of Africa. Furthermore, the agriculture sector in Africa is dominated by smallholder farmers with limited access to technology and the resources to adapt.[18]

Climate variability and change have been and continue to be, the principal source of fluctuations in global food production in countries of the developing world where production is highly rain-dependent.[19] The agriculture sector is sensitive to climate variability,[20] especially the inter-annual variability of precipitation, temperature patterns, and extreme weather events (droughts and floods). These climatic events are predicted to increase in the future and are expected to have significant consequences to the agriculture sector.[21] This would have a negative influence on food prices, food security, and land-use decisions.[22] Yields from rainfed agriculture in some African countries could be reduced by up to 50% by 2020.[21] In order to prevent the future destructive impact of climate variability on food production, it is crucial to adjust or suggest possible policies to cope with increased climate variability. African countries need to build a national legal framework to manage food resources in accordance with the anticipated climate variability. However, before devising a policy to cope with the impacts of climate variability, especially to the agriculture sector, it is critical to have a clear understanding of how climate variability affects different food crops.

In the year 2020, sever invasion of Locusts harmed the agriculture in eastern Africa. The invasion is at least partly due to climate change - the warmer temperature and heavier rainfall caused an abnormal increase in the number of locusts[23]

Water resources

Water quality and availability have deteriorated in most areas of Africa particularly due to climate change. Previous research and climate projections provide enough evidence that water resources are vulnerable and have the possibility of being strongly impacted by climate change with vast ramifications on human societies.[24]  The IPCC predicts millions of people in Africa will persistently face increased water stress due to climate variability and change (IPCC 2013). Changes in precipitation patterns directly affect surface runoff and water availability.[25] Any changes to the  hydrological cycle may have significant effects on river basins of Africa. To improve understanding of past and future changes in water availability due to climate change, the IPCC (IPCC 2013) recommends using the dynamic downscaling technique. The IPCC 2013 proposed using the coordinated regional downscaling experiment (CORDEX) regional climate models which runs at a maximum of 50  km resolutions, the resolution used depends upon the size of the watershed and area coverage by the meteorological records. However, before using the climate simulations from the dynamic downscaling, it is appropriate to evaluate their performance at different spatial scales since their performance differs from one location to another and from one RCM to another.

Health

African countries have the least efficient public health systems in the world.[26] Infectious disease burdens such as malaria, schistosomiasis, dengue fever, meningitis, which are sensitive to climate impacts, are highest in the sub-Saharan African region. For instance, over 90 percent of annual global malaria cases are in Africa.[26] Changes in climate will affect the spread of infectious agents as well as alter people’s disposition to these infections.

Energy

With increasing population and corresponding energy demand, energy security must be addressed because energy is crucial for sustainable development. Climate change has affected energy sectors in Africa as many countries depend on hydropower generation. Decreasing rainfall levels and droughts have resulted in lower water levels in dams with adverse impacts on hydropower generation. This has resulted in low electrical energy production, high cost of electricity and power outages or load-shedding in some African countries that depend on hydroelectric power generation. Disruptions in hydropower generation have negatively affected various sectors in countries such as Ghana, Uganda, Kenya, Tanzania.

Regional differences

Central Africa

Central Africa, for the most part, is landlocked and is geographically threatened by climate change. Due to its high climate variability and rainfed agriculture, Central Africa is expected to experience longer and more frequent heatwaves as well as an increase in wet extremes.[27] The global mean temperature in this region is to increase by 1.5 °C to 2 °C.[28]

Central Africa Adaption Measures

Angola - "The objective of the National Adaptation Programs of Action are to identify and communicate the urgent and immediate needs of the country regarding climate change adaptation, to increase Angola‘s resilience to climate variabilities and to climate change to ensure achievement of Poverty reduction programs, sustainable development objectives and the Millennium Development Goals pursued by the Government."[29]

Effects

The carbon dioxide-absorbing capacity of forests in the Congo Basin have decreased. This decrease has occurred due to increasing heat and drought causing decreased tree growth. This suggests that even unlogged forests are being affected by climate change. A Nature study indicates that by 2030, the African jungle will absorb 14 percent less carbon dioxide than it did from around 2005-2010, and will absorb none at all by 2035.[30]

Eastern Africa

Situated almost entirely in the tropics, rainfall in Eastern Africa is dominated by the seasonal migration of the tropical-rain band.[31] Eastern Africa is characterized by high spatio-temporal rainfall variability as it spans over 30 degrees of latitude (across the equator). It has influences from both the Indian and Atlantic Oceans, and has major orographic features (highlands) as well as  inland water bodies such as Lake Victoria. Therefore the rainfall seasonality varies from a single wet season per year in July–August in parts of the northwest (including Ethiopia and South Sudan, which are meteorologically more connected to West Africa, with the West African monsoon bringing the rains) to a single wet season per year in December - February in the south (over Tanzania), with many areas close to the equator having two rainy seasons per year,[32] approximately in March–May (the “Long Rains”) and October to December (the “Short Rains”). Fine-scale variability in rainfall seasonality is often linked to orography and lakes. Inter-annual variability can be large and known controls include variations in Sea surface temperatures (SSTs) of different ocean basins, large-scale atmospheric modes of variability such as the Madden-Julian Osciliation (MJO) [33][34] and tropical cyclones.[34][35]

Eastern Africa has witnessed frequent and severe droughts in recent decades, as well as devastating floods. Trends in rainfall since the 1980s show a general decrease in March - May (MAM) seasonal rains with a slight increase during June - September (JJAS) and October - December (OND) rains,[36] although there appears to have been a recent recovery in the MAM rains.[37] In the future, both rainfall and temperature are projected to change over Eastern Africa.[38][39][40] Recent studies on climate projections suggest that average temperature might increase by about 2-3 °C by the middle of the century and 2-5 °C at the end of the century.[41] This will depend on emission scenarios as well as on how the real climate responds compared with the range of possible outcomes shown by models. Climate model projections tend to show an increase in rainfall, particularly during OND season, which is also projected to occur later. This delay in the short rain season, has been linked to the deepening of the Saharan Heat Low under climate change.[42] It should be noted, however that some models predict decreasing rainfall,[40][41] and for some regions and seasons the very largest rainfall increases predicted have been shown to involve implausible mechanisms due to systematic model errors.[43]  In addition, changes of aerosols provide a forcing of rainfall change that is not captured in many assessments of climate projections.[44][45]

The contrast of the drying trend of MAM (long rains) rainfall in equatorial Eastern Africa, with most models predicting a wetting in the future has been labelled the “East African climate change paradox,"[45] although there has been some recent recovery in the rainfall.[37] Studies have shown that the drying trend is unlikely to be purely natural, but may be driven by factors such as aerosols rather than greenhouse gases,[45] further research is needed. The drying has been shown to be have been caused by a shorter rainy season, and linked to deepening of the Arabian Heat Low.[37]

Consistent with the uncertainty in rainfall projections, changes in rainy seasons onset are uncertain in equatorial Eastern Africa, although many models predict a later and wetter short rains.[42] The Indian Ocean Dipole (IOD) is known to provide a strong control on inter-annual variability in the short rains,[46] and studies show that extreme IODs may increase under climate change.[47]

Globally, climate change is expected to lead to intensification of rainfall, as extreme rainfall increases at a faster rate with warming than total rainfall does.[6] Recent work shows that across Africa global models are expected to under-estimate the rate of change of this rainfall intensification,[48] and changes in rainfall extremes may be much more widespread than those predicted by global models.[49]

Southern parts of Eastern Africa receive most of their rainfall in a single rainy season during the southern hemisphere’s winter: over Tanzania seasonal rainfall is projected to increase under future climate change, although there is uncertainty.[42] Further south, over Mozambique, a shorter season due to a later onset is projected under future climate change, again with some uncertainty.[50]

East Africa Adaption Measures

Comoros - "NAPA is the operational extension of the Poverty Reduction Strategy Paper (PRSP), as it includes among its adaptation priorities, agriculture, fishing, water, housing, health, but also tourism, in an indirect way, through the reconstitution of basin slopes and the fight against soils erosion, and therefore the protection of reefs by limiting the silting up by terrigenous contributions.[51]"

Kenya gazetted the Climate Change Act, 2016 which establishes an authority to oversee development, management, implementation and regulation of mechanisms to enhance climate change resilience and low carbon development for sustainable development, by the National and County Governments, the private sector, civil society, and other actors. Kenya has also developed the National Climate Change Action Plan (NCCAP 2018-2022) which aims to further the country's development goals by providing mechanisms and measures to achieve low carbon climate-resilient development in a manner that prioritizes adaptation.

Madagascar - the priority sectors for adaptation are: agriculture and livestock, forestry, public health, water resources and coastal zones.[52]

Malawi - The NAPA identifies the following as high priority activities for adaptation: "Improving community resilience to climate change through the development of sustainable rural livelihoods, Restoring forests in the Upper and Lower Shire Valleys catchments to reduce siltation and associated water flow problems, Improving agricultural production under erratic rains and changing climatic conditions, Improving Malawi’s preparedness to cope with droughts and floods, and Improving climate monitoring to enhance Malawi’s early warning capability and decision making and sustainable utilization of Lake Malawi and lakeshore areas resources[53]".

Mauritius - adaptation should address the following priority areas: coastal resources, agriculture, water resources, fisheries, health and well-being, land use change and forestry and biodiversity.[54]

Mozambique - "The proposed adaptation initiatives target various areas of economic and social development, and outline projects related to the reduction of impacts to natural disasters, the creation of adaptation measures to climate change, fight against soil erosion in areas of high desertification and coastal zones, reforestation and the management of water resources.[55]"

Rwanda has developed the National Adaptation Programme of Action (NAPA 2006) which contains information to guide national policy-makers and planners on priority vulnerabilities and adaptations in important economic sectors.[56] The country has also developed sector based policies on adaptation to climate change such as the Vision 2020, the National Environmental Policy and the Agricultural Policy among others.[57]

Tanzania Tanzania has outlined priority adaptation measures in their NAPA, and various national sector strategies and research outputs.[58] The NAPA has been successful at encouraging climate change mainstreaming into sector policies in Tanzania; however, the cross-sectoral collaboration crucial to implementing adaptation strategies remains limited due to institutional challenges such as power imbalances, budget constraints and an ingrained sectoral approach.[59] Most of the projects in Tanzania concern agriculture and water resource management (irrigation, water saving, rainwater collection); however, energy and tourism also play an important role.[60]

Zambia - "The NAPA identifies 39 urgent adaptation needs and 10 priority areas within the sectors of agriculture and food security (livestock, fisheries and crops), energy and water, human health, natural resources and wildlife[61]"

Zimbabwe - "The other strategic interventions by the NAP process will be: Strengthening the role of private sector in adaptation planning, Enhancing of the capacity of Government to develop bankable projects through trainings, Improving management of background climate information to inform climate change planning, Crafting a proactive resource-mobilization strategy for identifying and applying for international climate finance as requests for funds are primarily reactive at present, focusing on emergency relief rather than climate change risk reduction, preparedness and adaptation, Developing a coordinated monitoring and evaluation policy for programs and projects, as many institutions within the government do not currently have a systematic approach to monitoring and evaluation.[62] "

North Africa

Southern Africa

Southern Africa Adaption Measures

Lesotho - "The key objectives of the NAPA process entail: identification of communities and livelihoods most vulnerable to climate change, generating a list of activities that would form a core of the national adaptation program of action, and to communicate the country’s immediate and urgent needs and priorities for building capacity for adaptation to climate change[63]"

Namibia - the critical themes for adaptation are "Food security and sustainable biological resource base, Sustainable water resources base,Human health and well being and Infrastructure development[64]"

South Africa is in the progress of finalizing its national climate change adaptation strategy. "The National Adaptation strategy acts as a common reference point for climate change adaptation efforts in South Africa, and it provides a platform upon which national climate change adaptation objectives for the country can be articulated so as to provide overarching guidance to all sectors of the economy[65]"

West Africa and the Sahel

The West African region can be divided into four climatic sub-regions namely the Guinea Coast, Soudano-Sahel, Sahel (extending eastward to the Ethiopian border) and the Sahara,[66] each with different climatic conditions. The seasonal cycle of rainfall is mainly driven by the south-north movement of the Inter-Tropical Convergence Zone (ITCZ) which is characterised by the confluence  between moist southwesterly monsoon winds and the dry northeasterly Harmattan.[67]

Based on the inter-annual rainfall variability, three main climatic periods have been observed over the Sahel: the wet period from 1950 to the early 1960s followed by a dry period from 1972 to 1990 and then the period from 1991 onwards which has seen a partial rainfall recovery.[68][69][70] During the dry period, the Sahel experienced a number of particularly severe drought events, with devastating effects.[71][72] The recent decades, have also witnessed a moderate increment in annual rainfall since the beginning of 1990s. However, total annual rainfall remains significantly below that observed during the 1950s.[73][71]

Some have identified the recent 2 decades as a recovery period.[74] Others refer to this as a period of ‘hydrological intensification’ with much of the annual rainfall increase coming from more severe rain events and sometimes flooding rather than more frequent rainfall, or similarly other works [75][76] underline the continuity of the drought even though the rainfall has increased. Since 1985, 54 percent of the population has been affected by five or more floods in the 17 Sahel region countries.[77] In 2012, severe drought conditions in the Sahel were reported. Governments in the region responded quickly, launching strategies to address the issue.[78]

The region is projected to experience changes in rainfall regime, with climate models suggesting that decreases in wet season rainfall are more likely in the western Sahel, and increases more likely in the central to east Sahel, although opposite trends cannot yet be ruled out.[79][80][81] These trends will affect the frequency and severity of floods, droughts, desertification, sand and dust storms, desert locust plagues and water shortages.[82][83]

However, irrespective of the changes in seasonal mean rain, the most intense storms are expected to become more intense, amplifying flood frequency.[84][85] Enhanced carbon emissions and global warming may also lead to an increase in dry spells especially across the Guinea Coast associated with a reduction of the wet spells under both 1.5℃ and 2℃ global warming level.[86]

Fifteen percent of Sahel region population has also experienced a temperature increase of more than 1 °C from 1970 to 2010. The Sahel region, in particular, will experience higher average temperatures over the course of the 21st century and changes in rainfall patterns, according to the Intergovernmental Panel on Climate Change (IPCC).

African Highlands

Shifts in Malaria Transmission Due to Climate Change

Climate change, and resulting in increased temperatures, storms, droughts, and rising sea levels, will affect the incidence and distribution of infectious disease across the globe.[87] This is true in Africa, where malaria continues to have dramatic effects on the population. As climate change continues, the specific areas likely to experience the year-round, high-risk transmission of malaria will shift from coastal West Africa to an area between the Democratic Republic of the Congo and Uganda, known as the African Highlands.[88]

Exposures

To understand the exposures that affect shifting malaria transmission rates we can look to The Epidemiologic Triad, a model that explains the relationship between exposure, transmission, and causation of infectious diseases.[89] With regards to malaria transmission rates in the African Highlands, factors and exposures resulting from drastic environmental changes like warmer climates, shifts in weather patterns, and increases in human impact such as deforestation, provide appropriate conditions for malaria transmission between carrier and host.[90] Because of this, vectors will adapt, thrive, and multiply at a fast pace. An increase in the number of vectors that carry parasites, microbes, and pathogens that cause disease will become a health hazard for the human population.[87] Specifically, malaria is caused by the Plasmodium falciparum and Plasmodium vivax parasites which are carried by the vector Anopheles mosquito. Even though the Plasmodium vivax parasite can survive in lower temperatures, the Plasmodium falciparum parasite will only survive and replicate in the mosquito when climate temperatures are above 20℃.[91] Increases in humidity and rain also contribute to the replication and survival of this infectious agent.,[92] Increasing global temperatures combined with changes in land cover as a result of extreme deforestation will create ideal habitats for mosquitoes to survive in the African Highlands. If deforestation continues at its current rate, more land will be available for mosquito breeding grounds, and the population of mosquitos will rapidly increase. The increase in mosquitoes will thus increase the opportunity for both Plasmodium falciparum and Plasmodium vivax parasites to proliferate.

Exposure to malaria will become a greater risk to humans as the number of female Anopheles mosquitos infected with either the Plasmodium falciparum or Plasmodium vivax parasite increases.[92] The mosquito will transmit the parasite to the human host through a bite, resulting in infection. Then, when an uninfected mosquito bites the now infected human host, the parasite will be transmitted to the mosquitoes which will then become an exposure to other uninfected human hosts. Individuals who are constantly exposed to the Malaria parasite due to multiple bites by mosquitoes that carry the parasite are at greater risk of dying.[91] Infected humans can also transmit the disease to uninfected or healthy humans via contaminated blood.[91]

Health Effects

The health effects caused by shifts in malaria transmission rates in the African Highlands have the potential to be severe. Research has shown that the effects of climate change on health will impact most populations over the next few decades.[93] However, Africa, and specifically the African Highlands, are susceptible to being particularly negatively affected. In 2010, 91% of the global burden due to malaria deaths occurred in Africa. Several spatiotemporal models have been studied to assess the potential effect of projected climate scenarios on malaria transmission in Africa. A study conducted by Caminade et al.[94] concluded that the most significant climate change effects are confined to specific regions, including the African Highlands. These results are consistent with previous studies.

Ultimately, studies show an overall increase in climate suitability for malaria transmission resulting in an increase in the population at risk of contracting the disease.[94] Of significant importance is the increase of epidemic potential at higher altitudes (like the African Highlands). Rising temperatures in these areas have the potential to change normally non-malarial areas to areas with seasonal epidemics.[95] Consequently, new populations will be exposed to the disease resulting in healthy years lost. In addition, the disease burden may be more detrimental to areas that lack the ability and resources to effectively respond to such challenges and stresses.[96]

Scientific Limitations

Scientific limitations when examining shifting malaria transmission rates in the African Highlands are similar to those related to broader understandings of climate change and malaria. While modeling with temperature changes shows that there is a relationship between an increase in temperature and an increase in malaria transmission, limitations still exist. Future population shifts that affect population density, as well as changes in the behavior of mosquitos, can affect transmission rates and are limiting factors in determining the future risk of malaria outbreaks, which also affect planning for correct outbreak response preparation.[88]

Challenges and Solutions

The challenges of controlling and possibly eradicating malaria in the African Highlands are many and varied. Many of the strategies used to control malaria have not changed, are few in number and have rarely been added to in the last 20 years.

The most common forms of control are educating the public and vector control. The huge geographic area of the vectors Anopheles is possibly the largest challenge faced in the control of malaria. With such a large area to cover it is hard to use insecticides at a continuous and effective level.[97] This form of control is expensive, and the areas affected are not able to sustain control. Without sustained control, a rapid resurgence in parasite transmission is seen. Another challenge with insecticides is that the vector is now becoming insecticide-resistant. Due to the fact that mosquitoes have several generations per year, resistance is seen very quickly.[97]

Education has its limitations as well, as the population most affected by malaria are children, and the educational message is to stay inside during peak mosquito activity. The low socioeconomic status of the people who inhabit the African Highlands is also a challenge. Local health facilities have limited resources, and poor living conditions and malnourishment exacerbate malaria symptoms and increase the likelihood of death due to malaria.[97] As climate change shifts geographic areas of transmission to the African Highlands, the challenge will be to find and control the vector in areas that have not seen it before, and to not waste resources on areas where the temperature is no longer conducive to parasite growth.[98]

The solutions that can help malaria control and possibly lead to eradication are far fewer in number than the challenges, but if they are effective they can truly change the areas currently affected. There are number of groups working on a vaccine, some are looking to control the transmission of the parasite to the host, or control transmission from human back to the vector.[99] These vaccines are not very effective currently, and lose their effectiveness over time, so are not ideal. But, the development is still progressing in the hopes of finding a better, more effective long-lasting vaccine.[99] An alternative to vaccines is vectored immunoprophylaxis (VIP) that is a form a gene therapy. This therapy will change cells in the host that will secrete antigens from various stages of the parasite in the hopes of triggering an anamnestic immune response in the recipient and prevent disease and parasite transmission.[100]

Policy Implications

The policy implications of climate change and malaria rates in the African Highlands are also vast, and ultimately fall into two categories:

  1. Enacting policy that will reduce greenhouse gas emissions, thus slowing down climate change, and
  2. Mitigating problems that have already arisen, and will inevitably continue to develop, due to climate change.[101]

Addressing both of these areas is of great importance, as those in the poorest countries, including countries that make up the African Highlands, face the greatest burden. Additionally, when countries are forced to contend with a disease like malaria, their prospects for economic growth are slowed. This contributes to continued and worsening global inequality.[102]

When addressing policy that will reduce greenhouse gas emissions, it is necessary to act on a global scale, even when related effects are narrowed to a smaller area. The 2015 Lancet Commission on Health and Climate Change made nine recommendations for governments to address. These include:

  1. Make an investment in climate change research.
  2. Increase financing for global health systems.
  3. Eliminate coal as an energy source.
  4. Support cities that encourage healthy activities for individuals and the planet.
  5. Clarify carbon pricing.
  6. Increase access to renewable energy in low to middle-income countries.
  7. Quantify avoided burdens when these measures are taken.
  8. Collaborate with global governments and health organizations.
  9. Create an agreement that will help counties making changes to become low-carbon economies.[101]

When one focuses on mitigation, specifically as it relates to malaria in the African Highlands, research is still an important component. This research needs to take many forms, including attribution studies, to help clarify the degree to which malaria rates are attributed to climate change; scenario modeling, which can help further our understanding of future climate change consequences on malaria rates; and examinations of intervention programs and techniques, to help our understanding of what appropriate responses are.[102] Surveillance and monitoring of malaria in populations in the African Highlands will also be important, to better understand disease.[101]

Beyond these research priorities, it is also important that we enact policies that will significantly increase investments in public health in the African Highlands. This achieves two goals, the first being better outcomes related to malaria in the affected area, and the second being an overall better health environment for populations.[101] It is also important to focus on “one-health approaches."[101] This means collaborating on an interdisciplinary level, across various geographic areas, to come up with workable solutions.

These policies can be seen in action in the World Health Organization's “Adaptation to Climate Change in Africa Plan of Action for the Health Sector 2012-2016."[103] This report “is intended to provide a comprehensive and evidence-based coordinated response of the health sector to climate change adaptation needs of African countries in order to support the commitments and priorities of African governments."[103] The action plan includes goals like scaling up public health activities, coordinating efforts on an international scale, strengthening partnerships and collaborative efforts, and promoting research on both the effects of climate change as well as effective measures taken in local communities to mitigate climate change consequences.[103]

See also

Climate change by country in Africa


References

  1. Schneider, S.H.; et al. (2007). "19.3.3 Regional vulnerabilities". In Parry, M.L.; et al. (eds.). Chapter 19: Assessing Key Vulnerabilities and the Risk from Climate Change. Climate change 2007: impacts, adaptation, and vulnerability: contribution of Working Group II to the fourth assessment report of the Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press (CUP): Cambridge, UK: Print version: CUP. This version: IPCC website. ISBN 978-0-521-88010-7. Archived from the original on 2013-03-12. Retrieved 2011-09-15.
  2. Niang, I., O.C. Ruppel, M.A. Abdrabo, A. Essel, C. Lennard, J. Padgham, and P. Urquhart, 2014: Africa. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Barros, V.R., C.B. Field, D.J. Dokken et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1199-1265. https://www.ipcc.ch/site/assets/uploads/2018/02/WGIIAR5-Chap22_FINAL.pdf
  3. Ofoegbu, Chidiebere; Chirwa, P. W. (2019-05-19). "Analysis of rural people's attitude towards the management of tribal forests in South Africa". Journal of Sustainable Forestry. 38 (4): 396–411. doi:10.1080/10549811.2018.1554495. ISSN 1054-9811.
  4. Niang, I., O.C. Ruppel, M.A. Abdrabo, A. Essel, C. Lennard, J. Padgham, and P. Urquhart, 2014: Africa. In: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Barros, V.R., C.B. Field, D.J. Dokken et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom, and New York, NY, USA, pp. 1199-1265. https://www.ipcc.ch/site/assets/uploads/2018/02/WGIIAR5-Chap22_FINAL.pdf
  5. Ofoegbu, Chidiebere; Chirwa, P. W.; Francis, J.; Babalola, F. D. (2019-07-03). "Assessing local-level forest use and management capacity as a climate-change adaptation strategy in Vhembe district of South Africa". Climate and Development. 11 (6): 501–512. doi:10.1080/17565529.2018.1447904. hdl:2263/64496. ISSN 1756-5529.
  6. "Global Warming of 1.5 ºC —". Retrieved 2020-02-16.
  7. Rural societies in the face of climatic and environmental changes in West Africa. Impr. Jouve). Marseille: IRD éditions. 2017. ISBN 978-2-7099-2424-5. OCLC 1034784045.CS1 maint: others (link)
  8. Collins, Jennifer M. (2011-03-18). "Temperature Variability over Africa". Journal of Climate. 24 (14): 3649–3666. Bibcode:2011JCli...24.3649C. doi:10.1175/2011JCLI3753.1. ISSN 0894-8755. S2CID 129446962.
  9. Conway, Declan; Persechino, Aurelie; Ardoin-Bardin, Sandra; Hamandawana, Hamisai; Dieulin, Claudine; Mahé, Gil (2009-02-01). "Rainfall and Water Resources Variability in Sub-Saharan Africa during the Twentieth Century". Journal of Hydrometeorology. 10 (1): 41–59. Bibcode:2009JHyMe..10...41C. doi:10.1175/2008JHM1004.1. ISSN 1525-755X.
  10. "Southern African Development Community :: Climate Change Adaptation". www.sadc.int. Retrieved 8 August 2019.
  11. Lesolle, D (2012). SADC policy paper on climate change: Assessing the policy options for SADC member stated (PDF).
  12. Climate change adaptation in SADC: A Strategy for the water sector (PDF).
  13. "Programme on Climate Change Adaptation and Mitigation in Eastern and Southern Africa (COMESA-EAC-SADC)". Southern African Development Community. Retrieved 8 August 2019.
  14. AFRICAN STRATEGY ON CLIMATE CHANGE. https://www.un.org/en/africa/osaa/pdf/au/cap_draft_auclimatestrategy_2015.pdf: African Union. 2014.CS1 maint: location (link)
  15. Xu, Chi; M. Lenton, Timothy; Svenning, Jens-Christian; Scheffer, Marten (26 May 2020). "Future of the human climate niche (figure 3)". Proceedings of the National Academy of Sciences of the United States of America. 117 (21): 11350–11355. doi:10.1073/pnas.1910114117. PMID 32366654. Retrieved 4 June 2020.
  16. Supplementary Materials Future of the human climate niche (PDF). p. 21. Retrieved 14 June 2020.
  17. Läderach, Peter; Martinez-Valle, Armando; Bourgoin, Clement; Parker, Louis (2019-03-27). "Vulnerability of the agricultural sector to climate change: The development of a pan-tropical Climate Risk Vulnerability Assessment to inform sub-national decision making". PLOS ONE. 14 (3): e0213641. Bibcode:2019PLoSO..1413641P. doi:10.1371/journal.pone.0213641. ISSN 1932-6203. PMC 6436735. PMID 30917146.
  18. "Supporting Sub-Saharan Africa's Farmers - Bayer - Crop Science". www.cropscience.bayer.com. Retrieved 2019-08-15.
  19. Thornton, Philip K; Ericksen, Polly J; Herrero, Mario; Challinor, Andrew J (November 2014). "Climate variability and vulnerability to climate change: a review". Global Change Biology. 20 (11): 3313–3328. Bibcode:2014GCBio..20.3313T. doi:10.1111/gcb.12581. ISSN 1354-1013. PMC 4258067. PMID 24668802.
  20. "Empirical Relationships between Banana Yields and Climate Variability over Uganda". ResearchGate. Retrieved 2019-08-15.
  21. "Adapting Agriculture to Climate Change: Suitability of Banana Crop Production to Future Climate Change Over Uganda". ResearchGate. Retrieved 2019-08-15.
  22. "A1 - 1 Sustainability, food security and climate change: three intertwined challenges | Climate-Smart Agriculture Sourcebook | Food and Agriculture Organization of the United Nations". www.fao.org. Retrieved 2019-08-15.
  23. Rosane, Olivia (27 January 2020). "Worst Locust Swarm to Hit East Africa in Decades Linked to Climate Crisis". Ecowatch. Retrieved 6 February 2020.
  24. "Climate Change and Water — IPCC". Retrieved 2019-08-08.
  25. Fowler, H. J.; Blenkinsop, S.; Tebaldi, C. (October 2007). "Linking climate change modelling to impacts studies: recent advances in downscaling techniques for hydrological modelling". International Journal of Climatology. 27 (12): 1547–1578. Bibcode:2007IJCli..27.1547F. doi:10.1002/joc.1556. S2CID 53472608.
  26. World Health Organization. (2014). The health of the people: what works: the African Regional Health Report 2014. World Health Organization.
  27. Diedhiou, Arona; Bichet, Adeline; Wartenburger, Richard; Seneviratne, Sonia I.; Rowell, David P.; Sylla, Mouhamadou B.; Diallo, Ismaila; Todzo, Stella; Touré, N'Datchoh E.; Camara, Moctar; Ngatchah, Benjamin Ngounou; Kane, Ndjido A.; Tall, Laure; Affholder, François (2018). "Changes in climate extremes over West and Central Africa at 1.5 °C and 2 °C global warming". Environmental Research Letters. 13 (6): 065020. doi:10.1088/1748-9326/aac3e5.
  28. Seneviratne S I, Donat MG, Pitman AJ, Knutti R and Wilby R L 2016 Allowable CO2 emissions based on regional and impact-related climate targets Nature 529 477–83
  29. "analysis-of-human-rights-language-in-the-cancun-agreements-unfccc-16th-session-of-the-conference-of-the-parties". doi:10.1163/2210-7975_hrd-9986-0022. Cite journal requires |journal= (help)
  30. Grossman, Daniel (4 March 2020). "The Congo rainforest is losing ability to absorb carbon dioxide. That's bad for climate change". Washington Post. Retrieved 6 March 2020.
  31. Nicholson, Sharon E. (2017). "Climate and climatic variability of rainfall over eastern Africa". Reviews of Geophysics (in French). 55 (3): 590–635. Bibcode:2017RvGeo..55..590N. doi:10.1002/2016RG000544. ISSN 1944-9208.
  32. Dunning, Caroline M.; Black, Emily C. L.; Allan, Richard P. (2016). "The onset and cessation of seasonal rainfall over Africa". Journal of Geophysical Research: Atmospheres. 121 (19): 11, 405–11, 424. Bibcode:2016JGRD..12111405D. doi:10.1002/2016JD025428. ISSN 2169-8996.
  33. Vellinga, Michael; Milton, Sean (2018). "Drivers of interannual variability of the East African "Long Rains"". Quarterly Journal of the Royal Meteorological Society. 144 (1): 861–876. Bibcode:2018QJRMS.144..861V. doi:10.1002/qj.3263. ISSN 1477-870X.
  34. Finney, Declan L.; Marsham, John H.; Jackson, Lawrence S.; Kendon, Elizabeth J.; Rowell, David P.; Boorman, Penelope M.; Keane, Richard J.; Stratton, Rachel A.; Senior, Catherine A. (2019-02-05). "Implications of Improved Representation of Convection for the East Africa Water Budget Using a Convection-Permitting Model". Journal of Climate. 32 (7): 2109–2129. Bibcode:2019JCli...32.2109F. doi:10.1175/JCLI-D-18-0387.1. ISSN 0894-8755.
  35. Kilavi, Mary; MacLeod, Dave; Ambani, Maurine; Robbins, Joanne; Dankers, Rutger; Graham, Richard; Titley, Helen; Salih, Abubakr A. M.; Todd, Martin C. (December 2018). "Extreme Rainfall and Flooding over Central Kenya Including Nairobi City during the Long-Rains Season 2018: Causes, Predictability, and Potential for Early Warning and Actions". Atmosphere. 9 (12): 472. Bibcode:2018Atmos...9..472K. doi:10.3390/atmos9120472.
  36. Bernhofer, Christian; Hülsmann, Stephan; Gebrechorkos, Solomon H. (2019-08-06). "Long-term trends in rainfall and temperature using high-resolution climate datasets in East Africa". Scientific Reports. 9 (1): 11376. Bibcode:2019NatSR...911376G. doi:10.1038/s41598-019-47933-8. ISSN 2045-2322. PMC 6684806. PMID 31388068.
  37. Wainwright, Caroline M.; Marsham, John H.; Keane, Richard J.; Rowell, David P.; Finney, Declan L.; Black, Emily; Allan, Richard P. (2019-09-12). "'Eastern African Paradox' rainfall decline due to shorter not less intense Long Rains". NPJ Climate and Atmospheric Science. 2 (1): 1–9. doi:10.1038/s41612-019-0091-7. ISSN 2397-3722.
  38. Dunning, Caroline M.; Black, Emily; Allan, Richard P. (2018-10-04). "Later Wet Seasons with More Intense Rainfall over Africa under Future Climate Change" (PDF). Journal of Climate. 31 (23): 9719–9738. Bibcode:2018JCli...31.9719D. doi:10.1175/JCLI-D-18-0102.1. ISSN 0894-8755.
  39. Muhati, Godwin Leslie; Olago, Daniel; Olaka, Lydia (2018-10-01). "Past and projected rainfall and temperature trends in a sub-humid Montane Forest in Northern Kenya based on the CMIP5 model ensemble". Global Ecology and Conservation. 16: e00469. doi:10.1016/j.gecco.2018.e00469. ISSN 2351-9894.
  40. Rowell, David P.; Senior, Catherine A.; Vellinga, Michael; Graham, Richard J. (2016-02-01). "Can climate projection uncertainty be constrained over Africa using metrics of contemporary performance?". Climatic Change. 134 (4): 621–633. Bibcode:2016ClCh..134..621R. doi:10.1007/s10584-015-1554-4. ISSN 1573-1480.
  41. Bornemann, F. Jorge; Rowell, David P.; Evans, Barbara; Lapworth, Dan J.; Lwiza, Kamazima; Macdonald, David M.J.; Marsham, John H.; Tesfaye, Kindie; Ascott, Matthew J.; Way, Celia (2019-10-01). "Future changes and uncertainty in decision-relevant measures of East African climate". Climatic Change. 156 (3): 365–384. Bibcode:2019ClCh..156..365B. doi:10.1007/s10584-019-02499-2. ISSN 1573-1480.
  42. Dunning, Caroline M.; Black, Emily; Allan, Richard P. (2018-10-04). "Later Wet Seasons with More Intense Rainfall over Africa under Future Climate Change" (PDF). Journal of Climate. 31 (23): 9719–9738. Bibcode:2018JCli...31.9719D. doi:10.1175/JCLI-D-18-0102.1. ISSN 0894-8755.
  43. Rowell, David P. (2019). "An Observational Constraint on CMIP5 Projections of the East African Long Rains and Southern Indian Ocean Warming". Geophysical Research Letters. 46 (11): 6050–6058. Bibcode:2019GeoRL..46.6050R. doi:10.1029/2019GL082847. ISSN 1944-8007.
  44. Scannell, Claire; Booth, Ben B. B.; Dunstone, Nick J.; Rowell, David P.; Bernie, Dan J.; Kasoar, Matthew; Voulgarakis, Apostolos; Wilcox, Laura J.; Acosta Navarro, Juan C.; Seland, Øyvind; Paynter, David J. (2019-09-13). "The Influence of Remote Aerosol Forcing from Industrialized Economies on the Future Evolution of East and West African Rainfall". Journal of Climate. 32 (23): 8335–8354. Bibcode:2019JCli...32.8335S. doi:10.1175/JCLI-D-18-0716.1. ISSN 0894-8755.
  45. Rowell, David P.; Booth, Ben B. B.; Nicholson, Sharon E.; Good, Peter (2015-10-07). "Reconciling Past and Future Rainfall Trends over East Africa". Journal of Climate. 28 (24): 9768–9788. Bibcode:2015JCli...28.9768R. doi:10.1175/JCLI-D-15-0140.1. ISSN 0894-8755.
  46. Black, Emily; Slingo, Julia; Sperber, Kenneth R. (2003-01-01). "An Observational Study of the Relationship between Excessively Strong Short Rains in Coastal East Africa and Indian Ocean SST". Monthly Weather Review. 131 (1): 74–94. Bibcode:2003MWRv..131...74B. doi:10.1175/1520-0493(2003)131<0074:AOSOTR>2.0.CO;2. ISSN 0027-0644.
  47. Cai, Wenju; Wang, Guojian; Gan, Bolan; Wu, Lixin; Santoso, Agus; Lin, Xiaopei; Chen, Zhaohui; Jia, Fan; Yamagata, Toshio (2018-04-12). "Stabilised frequency of extreme positive Indian Ocean Dipole under 1.5 °C warming". Nature Communications. 9 (1): 1419. Bibcode:2018NatCo...9.1419C. doi:10.1038/s41467-018-03789-6. ISSN 2041-1723. PMC 5897553. PMID 29650992.
  48. Kendon, Elizabeth J.; Stratton, Rachel A.; Tucker, Simon; Marsham, John H.; Berthou, Ségolène; Rowell, David P.; Senior, Catherine A. (2019-04-23). "Enhanced future changes in wet and dry extremes over Africa at convection-permitting scale". Nature Communications. 10 (1): 1794. Bibcode:2019NatCo..10.1794K. doi:10.1038/s41467-019-09776-9. ISSN 2041-1723. PMC 6478940. PMID 31015416.
  49. Finney, Declan L.; Marsham, John H.; Rowell, David P.; Kendon, Elizabeth J.; Tucker, Simon O.; Stratton, Rachel A.; Jackson, Lawrence S. (2020-01-22). "Effects of explicit convection on future projections of mesoscale circulations, rainfall and rainfall extremes over Eastern Africa". Journal of Climate. 33 (7): 2701–2718. doi:10.1175/JCLI-D-19-0328.1. ISSN 0894-8755.
  50. Wang, Bin; Liu, De Li; Waters, Cathy; Yu, Qiang (2018-10-02). "Quantifying sources of uncertainty in projected wheat yield changes under climate change in eastern Australia". Climatic Change. 151 (2): 259–273. Bibcode:2018ClCh..151..259W. doi:10.1007/s10584-018-2306-z. ISSN 0165-0009.
  51. "National Action Programme of Adaptation to climate change (NAPA)" (PDF).
  52. "Madagascar". LDC Climate Change.
  53. "REPUBLIC OF MALAWI. MALAWI'S NATIONAL ADAPTATION PROGRAMMES OF ACTION (NAPA) ." (PDF).
  54. "Mauritius | UNDP Climate Change Adaptation". www.adaptation-undp.org. Retrieved 8 August 2019.
  55. "Mozambique National Adaptation Programs of Action (NAPA) | Climate and Development Learning Platform". www.climatelearningplatform.org. Retrieved 2019-08-08.
  56. USAID. "Climate Change Adaptation in Rwanda" (PDF). USAID.
  57. Ochieng, Cosmas; Khaemba, Winnie; Mwaniki, Ruchathi; Kimotho, Stephen (September 2017). Climate Change Adaptation in Rwanda's Agricultural Sector: A Case Study from Kirehe District, Eastern Province. Defending the Voiceless: Climate and Environmental Justice in Africa.
  58. USAID. "Climate Change Adaptation in Tanzania" (PDF). USAID.
  59. Pardoe, Joanna; Conway, Declan; Namaganda, Emilinah; Vincent, Katharine; Dougill, Andrew J.; Kashaigili, Japhet J. (9 August 2018). "Climate change and the water–energy–food nexus: insights from policy and practice in Tanzania". Climate Policy. 18 (7): 863–877. doi:10.1080/14693062.2017.1386082. ISSN 1469-3062.
  60. "United Republic of Tanzania | UNDP Climate Change Adaptation". www.adaptation-undp.org. Retrieved 8 August 2019.
  61. "National Adaptation Programme of Action on Climate Change (Zambia) | The REDD Desk". theredddesk.org. Retrieved 2019-08-08.
  62. "National Adaptation Plan(NAP) Roadmap for Zimbabwe" (PDF).
  63. "LESOTHO'S NATIONAL ADAPTATION PROGRAMME OF ACTION (NAPA) ON CLIMATE CHANGE" (PDF).
  64. "Youth Files Climate Case with India's Environmental Court". doi:10.1163/9789004322714_cclc_2017-0228-011. Cite journal requires |journal= (help)
  65. "NATIONAL CLIMATE CHANGE ADAPTATION STRATEGY REPUBLIC OF SOUTH AFRICA" (PDF).
  66. Agyeman, Richard Yao Kuma; Quansah, Emmannuel; Lamptey, Benjamin; Annor, Thompson; Agyekum, Jacob (2018). "Evaluation of CMIP5 Global Climate Models over the Volta Basin: Precipitation". Advances in Meteorology. Retrieved 2019-08-08.
  67. Sultan, Benjamin; Janicot, Serge (2003-11-01). "The West African Monsoon Dynamics. Part II: The "Preonset" and "Onset" of the Summer Monsoon". Journal of Climate. 16 (21): 3407–3427. Bibcode:2003JCli...16.3407S. doi:10.1175/1520-0442(2003)016<3407:TWAMDP>2.0.CO;2. ISSN 0894-8755.
  68. Le Barbé, Luc; Lebel, Thierry; Tapsoba, Dominique (2002-01-01). "Rainfall Variability in West Africa during the Years 1950–90". Journal of Climate. 15 (2): 187–202. Bibcode:2002JCli...15..187L. doi:10.1175/1520-0442(2002)015<0187:RVIWAD>2.0.CO;2. ISSN 0894-8755.
  69. Rowell, D.P., 2003: The Impact of Mediterranean SSTs on the Sahelian Rainfall Season. J. Climate, 16, 849-862
  70. Nicholson, SE, Fink, AH and Funk, C, 2018: Assessing recovery and change in West Africa's rainfall regime from a 161‐year record. Int. J. Climatol., 38, 3770-3786
  71. Funk, Chris; Fink, Andreas H.; Nicholson, Sharon E. (2018-08-01). "Assessing recovery and change in West Africa's rainfall regime from a 161‐year record". International Journal of Climatology. 38 (10): 3770–3786. Bibcode:2018IJCli..38.3770N. doi:10.1002/joc.5530.
  72. Nicholson, Sharon E. (2013). "The West African Sahel: A Review of Recent Studies on the Rainfall Regime and Its Interannual Variability". International Scholarly Research Notices. Retrieved 2019-08-08.
  73. Panthou, G.; et al. (2018). "Rainfall intensification in tropical semi-arid regions: the Sahelian case". Environmental Research Letters. 13 (6): 064013. Bibcode:2018ERL....13f4013P. doi:10.1088/1748-9326/aac334.
  74. Sanogo, Souleymane; Fink, Andreas H.; Omotosho, Jerome A.; Ba, Abdramane; Redl, Robert; Ermert, Volker (2015). "Spatio-temporal characteristics of the recent rainfall recovery in West Africa". International Journal of Climatology. 35 (15): 4589–4605. Bibcode:2015IJCli..35.4589S. doi:10.1002/joc.4309. ISSN 1097-0088.
  75. L'HOTE, YANN; MAHE, GIL; SOME, BONAVENTURE (2003-06-01). "The 1990s rainfall in the Sahel: the third driest decade since the beginning of the century". Hydrological Sciences Journal. 48 (3): 493–496. doi:10.1623/hysj.48.3.493.45283. ISSN 0262-6667.
  76. Nicholson, S. E.; Some, B.; Kone, B. (2000-07-01). "An Analysis of Recent Rainfall Conditions in West Africa, Including the Rainy Seasons of the 1997 El Niño and the 1998 La Niña Years". Journal of Climate. 13 (14): 2628–2640. Bibcode:2000JCli...13.2628N. doi:10.1175/1520-0442(2000)013<2628:AAORRC>2.0.CO;2. ISSN 0894-8755.
  77. Livelihood Security Climate Change, Migration and Conflict in the Sahel 2011
  78. Fominyen, George. "Coming weeks critical to tackle Sahel hunger – U.N. humanitarian chief". Thomson Reuters Foundation. Archived from the original on 3 June 2012. Retrieved 10 June 2012.
  79. Rowell, D.P., Senior, C.A., Vellinga, M. and Graham, R.J., 2016: Can Climate Projection Uncertainty be Constrained over Africa Using Metrics of Contemporary Performance? Climatic Change, 134, 621-633
  80. Berthou, S., Rowell D. P., Kendon E.J., Roberts. M.J, Stratton R., Crook J. and Wilcox C., 2019: Improved climatological precipitation characteristics over West Africa at convection-permitting scales. Climate Dynamics, 53, 1991–2011
  81. Kendon, Elizabeth J.; Stratton, Rachel A.; Tucker, Simon; Marsham, John H.; Berthou, Ségolène; Rowell, David P.; Senior, Catherine A. (2019). "Enhanced future changes in wet and dry extremes over Africa at convection-permitting scale". Nature Communications. 10 (1): 1794. Bibcode:2019NatCo..10.1794K. doi:10.1038/s41467-019-09776-9. ISSN 2041-1723. PMC 6478940. PMID 31015416.
  82. "IPCC Sees Severe Climate Change Impacts on Africa". ABC Live. ABC Live. Retrieved 7 September 2016.
  83. Vogel, Coleen. "Why Africa is particularly vulnerable to climate change". The Conversation. Retrieved 2017-08-07.
  84. Berthou, Ségolène; Rowell, David P.; Kendon, Elizabeth J.; Roberts, Malcolm J.; Stratton, Rachel A.; Crook, Julia A.; Wilcox, Catherine (2019-04-12). "Improved climatological precipitation characteristics over West Africa at convection-permitting scales". Climate Dynamics. 53 (3–4): 1991–2011. Bibcode:2019ClDy...53.1991B. doi:10.1007/s00382-019-04759-4. ISSN 0930-7575.
  85. Kendon, Elizabeth J.; Stratton, Rachel A.; Tucker, Simon; Marsham, John H.; Berthou, Ségolène; Rowell, David P.; Senior, Catherine A. (2019-04-23). "Enhanced future changes in wet and dry extremes over Africa at convection-permitting scale". Nature Communications. 10 (1): 1794. Bibcode:2019NatCo..10.1794K. doi:10.1038/s41467-019-09776-9. ISSN 2041-1723. PMC 6478940. PMID 31015416.
  86. Klutse, Nana Ama Browne; Ajayi, Vincent O.; Gbobaniyi, Emiola Olabode; Egbebiyi, Temitope S.; Kouadio, Kouakou; Nkrumah, Francis; Quagraine, Kwesi Akumenyi; Olusegun, Christiana; Diasso, Ulrich (May 2018). "Potential impact of 1.5\hspace0.167em°C and 2\hspace0.167em°C global warming on consecutive dry and wet days over West Africa". Environmental Research Letters. 13 (5): 055013. doi:10.1088/1748-9326/aab37b. ISSN 1748-9326.
  87. Beard, C. B., Eisen, R. J., Barker, C. M., Garofalo, J. F., Hahn, M., Hayden, M., . . . Schramm, P. J. (2016). Vector-Borne Diseases. Retrieved February 15, 2017, from https://health2016.globalchange.gov/vectorborne-diseases
  88. Ryan, Sadie J.; McNally, Amy; Johnson, Leah R.; Mordecai, Erin A.; Ben-Horin, Tal; Paaijmans, Krijn; Lafferty, Kevin D. (2015). "Mapping Physiological Suitability Limits for Malaria in Africa Under Climate Change". Vector-Borne and Zoonotic Diseases. 15 (12): 718–725. doi:10.1089/vbz.2015.1822. PMC 4700390. PMID 26579951.
  89. Centers for Disease Control and Prevention. (2012). Lesson 1: Introduction to Epidemiology. Retrieved March 26, 2017, fromhttps://www.cdc.gov/ophss/csels/dsepd/ss1978/lesson1/section8.html
  90. Himeidan, Y. E.; Kweka, E. J. (2012). "Malaria in East African highlands during the past 30 years: Impact of environmental changes". Frontiers in Physiology. 3: 315. doi:10.3389/fphys.2012.00315. PMC 3429085. PMID 22934065.
  91. "Where Malaria Occurs". Center for Disease Control and Prevention. 14 November 2018. Retrieved 27 February 2020.
  92. "Climate Change and Vector-Borne Disease". Center for Science Education. 2011. Retrieved 27 February 2020.
  93. Costello, Anthony; Abbas, Mustafa; Allen, Adriana; Ball, Sarah; Bell, Sarah; Bellamy, Richard; Friel, Sharon; Groce, Nora; Johnson, Anne; Kett, Maria; Lee, Maria; Levy, Caren; Maslin, Mark; McCoy, David; McGuire, Bill; Montgomery, Hugh; Napier, David; Pagel, Christina; Patel, Jinesh; de Oliveira, Jose Antonio Puppim; Redclift, Nanneke; Rees, Hannah; Rogger, Daniel; Scott, Joanne; Stephenson, Judith; Twigg, John; Wolff, Jonathan; Patterson, Craig (May 2009). "Managing the health effects of climate change". The Lancet. 373 (9676): 1693–1733. doi:10.1016/S0140-6736(09)60935-1. PMID 19447250.
  94. Caminade, Cyril; Kovats, Sari; Rocklov, Joacim; Tompkins, Adrian M.; Morse, Andrew P.; Colón-González, Felipe J.; Stenlund, Hans; Martens, Pim; Lloyd, Simon J. (2014). "Impact of climate change on global malaria distribution". Proceedings of the National Academy of Sciences. 111 (9): 3286–3291. Bibcode:2014PNAS..111.3286C. doi:10.1073/pnas.1302089111. PMC 3948226. PMID 24596427.
  95. Martens, W. J., Niessen, L. W., Rotmans, J., Jetten, T. H., & McMichael, A. J. (1995). The potential impact of global climate change on malaria risk. Environmental Health Perspectives, 103(5), 458–464.
  96. Wu, Xiaoxu; Lu, Yongmei; Zhou, Sen; Chen, Lifan; Xu, Bing (2016). "Impact of climate change on human infectious diseases: Empirical evidence and human adaptation". Environment International. 86: 14–23. doi:10.1016/j.envint.2015.09.007. PMID 26479830.
  97. Osungbade, K. O., & Oladunjoye, O. O. (2012). Prevention of Congenital Transmission of Malaria in Sub- Saharan African Countries: Challenges and Implications for Health System Strengthening. Journal of Tropical Medicine, 2012, 1-6. doi:10.1155/2012/648456
  98. Tanser, F. C., Sharp, B., & Sueur, D. L. (2003). The potential effect of climate change on malaria transmission in Africa. The Lancet, 362(9398), 1792-1798. doi:10.1016/s0140-6736(03)14898-2
  99. Nunes, J. K., Woods, C., Carter, T., Raphael, T., Morin, M. J., Diallo, D., . . . Birkett, A. J. (2014). Development of a transmission-blocking malaria vaccine: Progress, challenges, and the path forward. Vaccine, 32(43), 5531-5539. doi:10.1016/j.vaccine.2014.07.030
  100. Rodrigues, M. M., & Soares, I. S. (2014). Gene-therapy for malaria prevention. Trends in Parasitology, 30(11), 511-513. doi:10.1016/j.pt.2014.09.005
  101. Watts, Nick; Adger, W Neil; Agnolucci, Paolo; Blackstock, Jason; Byass, Peter; Cai, Wenjia; Chaytor, Sarah; Colbourn, Tim; Collins, Mat; Cooper, Adam; Cox, Peter M.; Depledge, Joanna; Drummond, Paul; Ekins, Paul; Galaz, Victor; Grace, Delia; Graham, Hilary; Grubb, Michael; Haines, Andy; Hamilton, Ian; Hunter, Alasdair; Jiang, Xujia; Li, Moxuan; Kelman, Ilan; Liang, Lu; Lott, Melissa; Lowe, Robert; Luo, Yong; Mace, Georgina; et al. (2015). "Health and climate change: Policy responses to protect public health" (PDF). The Lancet. 386 (10006): 1861–1914. doi:10.1016/S0140-6736(15)60854-6. hdl:10871/17695. PMID 26111439.
  102. Campbell-Lendrum, D.; Manga, L.; Bagayoko, M.; Sommerfeld, J. (2015). "Climate change and vector-borne diseases: What are the implications for public health research and policy?". Philosophical Transactions of the Royal Society B: Biological Sciences. 370 (1665): 20130552. doi:10.1098/rstb.2013.0552. PMC 4342958. PMID 25688013.
  103. World Health Organization. (2012). Adaptation to climate change in Africa plan of action for the health sector 2012-2016. Retrieved from http://www.afro.who.int/index.php?option=com_docman&task=doc_download&gid=7699&Itemid=2593
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.