Impacts of climate change on the livestock food supply chain; a review of the evidence (2024)

As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsem*nt of, or agreement with, the contents by NLM or the National Institutes of Health.
Learn more: PMC Disclaimer | PMC Copyright Notice

Impacts of climate change on the livestock food supply chain; a review of the evidence (1)

Glob Food Sec. 2021 Mar; 28: 100488.

Published online 2021 Mar. doi:10.1016/j.gfs.2020.100488

PMCID: PMC7938222

PMID: 33738188

Author information Article notes Copyright and License information PMC Disclaimer

Associated Data

Supplementary Materials

Abstract

The potential impacts of climate change on current livestock systems worldwide are a major concern, and yet the topic is covered to a limited extent in global reports such as the ones produced by the Intergovernmental Panel on Climate Change. In this article, we review the risk of climate-related impacts along the land-based livestock food supply chain. Although a quantification of the net impacts of climate change on the livestock sector is beyond the reach of our current understanding, there is strong evidence that there will be impacts throughout the supply chain, from farm production to processing operations, storage, transport, retailing and human consumption. The risks of climate-related impacts are highly context-specific but expected to be higher in environments that are already hot and have limited socio-economic and institutional resources for adaptation. Large uncertainties remain as to climate futures and the exposure and responses of the interlinked human and natural systems to climatic changes over time. Consequently, adaptation choices will need to account for a wide range of possible futures, including those with low probability but large consequences.

Keywords: Livestock, Climate change, Supply chain, Heat stress, Vulnerability, Risk

Graphical abstract

Highlights

  • Risk results from the interaction of climate-related hazards with the exposure and vulnerability of human and natural systems.

  • Climate change will impact the livestock sector throughout the food supply chain⁠—from farm production to human consumption.

  • Key hazards relate to climate change trends but also, and importantly, to climate variability and climate extremes.

  • Large uncertainties remain as to climate futures and the exposure and responses of the interlinked human and natural systems.

  • Adaptation choices will need to account for a wide range of possible futures.

1. Introduction

Climate change is a major concern for current livestock systems worldwide. Global warming and its associated changes in mean climate variables and climate variability affect feed and water resources as well as animal health and production. Climate change also has implications for the processing, storage, transport, retailing and consumption of livestock products. The ability of current livestock systems to support livelihoods and meet the increasing demand for livestock products is thus threatened.

The livestock sector currently plays a key role in food supply and food security. Livestock products (meat, milk and eggs) contribute 15% and 31% of global per capita calorie and protein supply, with regional variations (FAOSTAT, 2020; see Appendix for calculation of estimates presented in the Introduction). About 30% and 6% of global ruminant meat and milk production originates from grazing systems, on land that is often poorly suited for cropping (Herrero et al., 2013). Furthermore, livestock provides a range of other services, including as a source of draught power, a means of transportation, a source of nutrients for poor soils, a source of income generation and diversification, and a form of financial capital, all of which contribute to the overall well-being and resilience of many communities (CIRAD, 2016). Over 844 million people worldwide receive some income from agriculture, and the livestock sector contributes about 40% of agricultural value-added (FAOSTAT, 2020; The World Bank, 2020a). Livestock contributions to food security and other sustainability dimensions will be affected by climate change, although the full extent and magnitude of the impacts remain unknown.

Livestock and climate change studies often focus on the climate mitigation potential of livestock and in describing adaptation practices. When studies cover climate impacts, these tend to have a relatively narrow viewpoint, focusing on specific livestock species, primary production, or on selected dimensions of risk of climate-related impacts such as climate hazards without considering vulnerability levels of different communities (e.g. of reviews, Escarcha et al. (2018); IPCC (2014); Rivera-Ferre et al. (2016); Rojas-Downing et al. (2017); Thornton et al. (2009)). In large part this reflects the fact that, compared to crop production, considerably less work has been published on observed and modelled climate impacts on livestock (IPCC, 2014a). It also reflects the limited number of synthetic reviews of the issue, as highlighted in Rivera-Ferre et al. (2016).

In order to fill this gap, we review the risk of climate-related impacts along the land-based livestock food supply chain (i.e. from production to consumption). While not exhaustive, we aim to capture the major trends with direct implications for livestock-sourced food availability, access, utilisation and stability, and highlight key recent literature. We acknowledge that the implications of climate change go well beyond these pillars and affect the provision of goods and services (e.g. wool, hides, skins and manure, animal traction, financial instrument, etc.), human livelihoods and health, ecosystems, economies, cultures, and infrastructure in complex ways. Also, while we recognize that climate adaptation strategies and the impacts of livestock on climate change are significant considerations, these are not covered here but assessed elsewhere (e.g. FAO (2018a), Escarcha et al. (2018), Henry et al. (2018), Herrero et al. (2016), Rivera-Ferre et al. (2016), Salman et al. (2019), Sejian et al. (2015) and Weindl et al. (2015)).

This review is framed around the concept of risk of climate-related impacts, as defined by the Intergovernmental Panel on Climate Change (IPCC) Working Group II (IPCC, 2014a). Risk of climate-related impacts results from the interaction of climate-related hazards with the exposure and vulnerability of human and natural systems (Fig. 1). The analysis of this interaction represents the core of the IPCC climate impacts assessments. We use the term hazard to refer to climate-related physical events or trends that impact livestock systems (IPCC 2014). Exposure refers to the parts of the livestock supply chain that could be adversely affected, while vulnerability encompasses humans’ capacity to cope and adapt to changes. The term impact is used primarily to refer to the effects of extreme weather, climate events and climate change on natural and human systems.

Impacts of climate change on the livestock food supply chain; a review of the evidence (3)

Schematic of the interaction among the physical climate system, exposure, and vulnerability producing risk in the livestock supply chain. Risk of climate-related impacts results from the interaction of climate-related hazards (including hazardous events and trends) with the vulnerability and exposure of human and natural systems. Changes in both the climate system (left) and socioeconomic processes including adaptation and mitigation (right) are drivers of hazards, exposure, and vulnerability. Adapted from IPCC (2014).

We first detail the extent to which the livestock supply chain is exposed to climate change, referring to key literature on topics that have received significant past attention and expanding on recent topics of concern. We then discuss the societal ability of the livestock sector to cope or adapt to changes considering broader societal trends before highlighting potential risks of climate-related impacts.

2. Exposure of human and natural systems

Anthropogenic greenhouse gas emissions are associated with a range of climate and atmospheric shifts, the impacts of which are already being observed (FAO, IFAD, UNICEF, 2018). Major trends that can impact the livestock food supply chain are increases in atmospheric carbon dioxide (eCO2) and tropospheric ozone (O3) concentrations; changes in both mean and variability of temperature and precipitation; sea level rise and storm surges; and increased risk and frequency of extreme weather events. Past and projected changes in these climate variables are detailed in the Supplementary Information.

These climate change hazards may adversely affect the livestock sector at different stages of the livestock supply chain, as summarised in Fig. 2 and Table 1, and further detailed in the sections below. The potential impacts on labour and prices and overall social dynamics, which affect all stages of the supply chain, are described last.

Impacts of climate change on the livestock food supply chain; a review of the evidence (4)

Potential impacts of climate-related hazards on the livestock land-based food supply chain. The term quantity encompasses here the notions of physical availability of feed and animal products, economic and physical access and stability of these products (availability, access and stability indicators as defined by FAO (2019)). The human livelihood capitals listed follow the sustainable livelihoods framework introduced by Scoones (1998) and thereafter modified (Ellis, 2000). Relevant text sections are provided in brackets.

Table 1

Summary of the potential impacts of climate variables on the livestock food supply chain. The nature, extent and magnitude of the impacts will vary depending on a range of factors not described here that may interact in complex ways, such as interactions between climate variables, plant and animal species, agro-ecological conditions and socio-economic contexts.

Supply chainPotential impacts of climate change
Feed resourcesProductivity Regions already water stressed are likely to experience the most negative impacts. Some regions in high latitudes could experience yield increases due to reduced cold stress and longer growing seasons. Soil salinity in coastal regions may increase due to sea level rise and increased frequency and intensity of storm surges. Changing precipitation patterns particularly in arid regions could contribute to greater salinity. Changing weather patterns and warming temperatures could contribute to shifting pest and disease distribution and could increase stress on key pollinator species. Hotter and more humid conditions are likely to result in increased on-farm post-harvest losses where storage conditions are inadequate.
Elevated eCO2 can increase yields but won't benefit all crops equally. Temperate C3 species could be the most positively affected, and realised benefits may be mitigated by water and nutrient constraints. Elevated O3 will have a negative effect on yields.
Nutritional quality Warmer temperatures and drier conditions will tend to favour C4 species and increase toxicity in some plants, including during storage. Elevated eCO2 could reduce plant protein and mineral concentrations and increase toxicity in some species. Increases in eCO2 will tend to favour C3 plants and woody encroachment at the expense of grasses.
Variability in feed availability Inter-annual climate variability is projected to increase globally with overall negative impact on feed production. Changes in seasonal climate patterns will have context specific impacts, which may be positive or negative. However, increased variability will likely lead to less predictable feed supply. Extreme events could restrict animal access to pastures and create larger disruptions to feed production.
Water resourcesHotter and drier conditions are likely to increase water requirements of plants and animals, increasing pressure on water resources, especially in regions already water stressed. Further, warming temperatures will contribute to greater glacier depletion disrupting historical surface water flows.
Higher temperatures and extreme events such as floods and droughts are likely to decrease water quality for animal consumption, through increased concentration of pathogens, sediments, salts, nutrients or pollutants in water.
Animal health and productionAnimal production, welfare and life expectancy are likely to be negatively impacted, through decreased feed availability and quality, heat stress, diseases (from outbreaks and weakened animal immune system) and mortality from extreme climate events such as storms, floods, heat and cold waves. Globally, the effects are likely to be negative, but in some geographies with cold winters, warmer temperatures may reduce animal cold stress and maintenance energy requirements, as well as housing heating.
Processing, storage, transport and retailingHigher temperatures, increased humidity, increased frequency of extreme weather events, and rising sea levels are likely to put additional stress on built-up capital (machinery, transportation infrastructure, electricity networks, telecommunications, etc.). Further, warmer temperatures could increase the risk of animal heat stress during transportation, as well as worsen conditions for storage and distribution of food and feed, which could lead to reduced food quality, safety and shelf-life.
Increased variability in production and extreme climate events will likely make trade patterns less regular, increasing reliance on complex logistic systems.
Livestock products consumptionClimate change can reduce the availability of livestock products, as well as their quality and safety through contamination with pathogens or pesticide and reduced nutritional quality and sensory appeal. Prices may increase and be more volatile. Changing social norms may impact diets, especially in high-income countries.
LabourLabour availability and productivity is likely to be negatively impacted by climate change due to heat stress, increased risk of novel disease outbreaks, and extreme events like heat waves, floods and severe storms. Labour is also likely to be negatively impacted by exposure to decreased air quality associated with rising temperatures, nutrition from changes in food supply.
PricesCosts along the supply chain, commodity price and price volatility are likely to increase under climate change. The impacts of climate change on animal product prices could be felt mainly through changes in costs and availability of feed.

2.1. Feed and water resources

2.1.1. Quantity and quality of livestock feed production

Changes to the quantity and quality of livestock feed will be influenced by complex local interactions between eCO2 concentrations, tropospheric O3 levels, temperature, and precipitation. We first provide an inventory of how eCO2, O3, temperature and precipitation can affect livestock feed, then present some model projections under climate change. Livestock consume grains (especially in poultry, pig and intensive ruminant systems), crop above-ground biomass (e.g. in dual purpose crops which are both grazed and harvested), crop residues (e.g., straw or stover – key feed in mixed crop-livestock systems) as well as native and sown pastures (key feed in mixed crop-livestock and grazing-only systems). While not covered here, livestock can also be fed by-products and waste (e.g. oilseed cakes, bran, vegetable waste, brewer waste), concentrates and supplements (FAO, 2017).

2.1.1.1. Direct impacts of atmospheric CO2 and tropospheric O3 on feed

Research shows that eCO2 may have both positive and negative impacts on livestock feed, although there is recent evidence that the fertilisation effects of eCO2 and nitrogen on plant physiological processes may slow in the future as ecosystems productivity become dominated by the negative effects of higher temperatures and extreme droughts (Peñuelas et al., 2017).

Increases in eCO2 concentrations stimulate plant primary productivity (see review in Ainsworth et al., 2020), increasing potential yields of some species. Plants with a C3 photosynthetic pathway such as wheat, rice, soybean and temperate grasses experience greater growth stimulation than C4 plants such as maize, sorghum, sugarcane and tropical grasses. However, the CO2 fertilisation effects can also reduce animal feed quality (Augustine et al., 2018; Myers et al., 2014; Smith and Myers, 2018). For example, Myers et al. (2014) reported that C3 crops other than legumes had lower grain protein concentrations under elevated eCO2 concentration in the range 546–586ppm (−6.3% in wheat grains and −7.5% in rice grains). The impact on C4 crop grain was smaller. Increased eCO2 was also found to decrease the overall mineral concentrations (−8%) and increased the total non-structural carbohydrate (mainly starch, sugars) to mineral ratios in the total biomass of non-leguminous C3 plants (Loladze, 2014). While the nutritional quality of C3 grasses may be the most greatly impacted by eCO2 increases, it may nonetheless remain higher than C4 grasses under elevated eCO2 (Barbehenn et al., 2004). Increased toxicity has also been reported in some plants, with Gleadow et al. (2009) measuring a 160% increase in the concentration of cyanogenic glycosides (compounds that break down to release toxic hydrogen cyanide when plant tissue is crushed or chewed) in cassava leaves between CO2 concentrations of 360 and 710ppm in greenhouse experiments. Woody encroachment associated with rising eCO2 levels and changes in fire regimes can also alter grassland ecosystem function and negatively impact the intake and quality of grazing animals’ diets. Woody forages are harder for cattle and sheep to physically access as compared to goats, are less palatable, and have lower dry matter and protein digestibility compared to herbaceous plants (Archer et al., 2017).

The effects of increasing tropospheric O3 on plant productivity at scale and the range of potential secondary effects it might have (e.g. on weeds, pests and diseases, interactions with chemicals such as pesticides) have received less attention than eCO2 (Ainsworth et al., 2020). However, synthesis of crop responses to O3 finds that O3 pollution reduces crop yields to a similar level as nutrient, heat and aridity stress (Mills et al., 2018). For instance, using historical ground-level monitoring data, McGrath et al. (2015) estimated that, over the past 30 years, O3 pollution reduced U.S. soybean and maize yields by 5–10%. Ainsworth et al. (2020) provide a review of the literature on the effect of this air pollutant on plant productivity. The topic is not yet fully understood and remains one of the key uncertainties in crop, grassland and other global terrestrial models, with significant implications on our ability to predict future atmospheric composition and global climate, net primary productivity and provision of ecosystem services.

2.1.1.2. Direct impacts of water and temperature on feed

Changes in temperature and water availability can greatly affect forage and crop yields and feed quality. Sensitivity to changes in climate depends on the crop type and other environmental factors, but there is strong agreement that air temperatures above approximately 30°C–34°C generally depress cereal yields under water-limited conditions, through accelerating crop development and damaging plant cells (Carlson, 1990; D. B. Lobell et al., 2011; Meerburg et al., 2009). The maximum temperature for growth of temperate legumes and pastures is around 30–35°C, increasing to 35–50°C for tropical species (Ludlow, 1980). High temperatures are often coupled with water stress, since low soil moisture results in a decrease in evaporative cooling from the landscape (Mueller and Seneviratne, 2012) and high temperatures increase crop water loss (Lobell et al., 2013). The combination of warmer temperatures and drier conditions tends to favour C4 rather than C3 species (Hatfield et al., 2011; Izaurralde et al., 2011). The concentration of ergot alkaloids, and other potentially toxic secondary compounds (e.g., hydrogen cyanide in cassava and forage sorghum) are also likely to increase in response to a hotter and drier climate (Bourguignon et al., 2015; Brown et al., 2016; Gleadow et al., 2016).

Increased instability of feed supply is particularly a concern in grazing systems where it represents a major challenge for herd size and grazing intensity management (Godde et al., 2020; Sayre et al., 2013; Sloat et al., 2018). Pastures with high year-to-year precipitation variability were found to currently support lower livestock stocking rates than less variable regions (Sloat et al., 2018). Studies focused on grassland vegetation have also found that changes in seasonal climate patterns can have either positive or negative impacts on above ground biomass, depending on the nature of the change and the agro-ecological context (Craine et al., 2012; Guan et al., 2014; Peng et al., 2013; Prevéy and Seastedt, 2014; von Wehrden et al., 2010; Zeppel et al., 2014). The arrangement of climate extreme sequences such as drought sequences or number of hot days in a row, could have significant implications for the livestock sector (Stafford Smith and Foran, 1992). While less commonly researched than droughts, other hazards such as fires, heavy storms, flooding events, surface melt and icing events, as well as the appearance of new lakes, streams and marshes also disturb crop growth, reduce arable land and restrict animal access to pastures (Amstislavski et al., 2013; Pan et al., 2019). For instance, in northern Russia, nomadic reindeer herders migrate hundreds of kilometres in spring and autumn to connect summer and winter pastures. The appearance of new water bodies and change in size of existing ones due to melting permafrost can act as barriers, changing migration routes and increasing grazing pressure on the most accessible pastures (Amstislavski et al., 2013).

Changes in precipitation patterns in saline areas will also affect soil salinity and agricultural production potentials. Salinity intrusion and associated reductions in forage area have led farmers across the coastal belt in Bangladesh to look for other sources of livestock feed (Alam et al., 2017). Tajul Baharuddin et al. (2013) suggest that the predicted local sea-level rise for areas such as Carey Island in Malaysia would prevent oil palm production by the 21st century, due to seawater intrusion. This has implications for livestock production through potential reductions in the production of palm kernel meal, which is often fed to cattle in industrial systems. Integrated palm-cattle systems, where cattle graze under trees or are fed palm fronds removed as part of plantation maintenance, would also be impacted. Increases in the frequency, duration and intensity of heavy rainfall events, drought periods and sea level rise will also increase exposure of water, croplands and grasslands to soil contaminants with potential harmful impacts for crop and forage yield quantity and quality (Biswas et al., 2018; Lemonte et al., 2017; Marrugo-Negrete et al., 2019).

2.1.1.3. Feed yields, as projected in the future by biophysical models

At higher levels of warming, crop yields are projected to drop, especially at lower latitudes (Rosenzweig et al., 2014). This is particularly the case for maize and wheat yields which begin to decline with 1°C–2°C of local warming in the tropics, and drop by up to 60% under 5°C of local warming (IPCC, 2014a). Temperate maize is less clearly affected at the 1–⁠2°C threshold, but would be significantly affected with warming of 3°C–5°C. Recent studies also show that global food production has likely already been impacted (Asseng et al., 2015; Lobell et al., 2011; Ray et al., 2019). Ray et al. (2019) estimated that the impact of observed climate change on yields of different crops ranged from −13.4% (oil palm) to +3.5% (soybean), with impacts mostly negative in Europe, Southern Africa and Australia but generally positive in Latin America. Crop yield interannual variability is likely to progressively increase in many regions (IPCC, 2014a). For instance, Müller and Robertson (2014), in a gridded modelling study reported an increase of interannual variability of more than 5% in 64% of grid cells, and a decrease of more than 5% in 29% of cases by 2050.

Regarding forage availability, as for food crops, the diversity and severity of likely impacts differ considerably by location and species. In an assessment of global rangelands, Godde et al. (2020) found that global mean herbaceous biomass is projected to decrease of 4.7% by 2050 under RCP 8.5, with 74% of global rangeland area projected to experience a decline in mean biomass. The largest regional decrease was projected for Oceania while the highest increase was found for Europe. Another study focussed on European grasslands (Chang et al., 2017) also found projected increases in grassland productivity, mainly attributed to the simulated fertilisation effect of rising CO2. Both studies highlight projected increases in biomass inter-annual variability over some regions. Woody encroachment is also projected to occur on over 51% of global rangeland area by 2050 under RCP 8.5 according to Godde et al. (2020).

In an integrated partial equilibrium modelling approach, Havlík et al. (2015) found that the climate change impacts on crop and pasture forage yields will have little effect on global milk and meat production by 2050 due to trade in animal products, which could compensate for the feed deficits in some parts of the world. However, depending on the scenario, the impacts could be more pronounced at the regional scale. The most uncertain and potentially the most severe effects were found for sub-Saharan Africa, where for example, ruminant meat production could increase by 20% by 2050 but could also decrease by 17%, depending on projected feed supply based on varying the biophysical crop model and CO2 fertilisation assumptions.

Yield projections as above-mentioned are subject to large uncertainties. Models do not usually consider extreme events and overall increasing variability, or explicitly represent adaptation or effects such as changes in tropospheric O3, pests, pollinators, or agricultural labour. There are also large uncertainties as to climate extremes and trends in the future (Eyring et al., 2019; Sillmann et al., 2017) and changes in management practices, historical and projected land-use patterns (Polley et al., 2017). Our understanding of ecosystems responses to climate change is also limited (Rosenzweig et al., 2014; Schewe et al., 2019), including the interacting consequences of changes in temperature, precipitation, eCO2 and tropospheric O3, particularly in the context where management-driven yields increases are still occurring across vast areas of croplands. Possible ecosystems transitions from equilibrium to non-equilibrium systems driven primarily by stochastic abiotic factors will likely result in highly variable and less predictable primary production. Land uses such as grazing can also regulate grasslands responses to climate change. For example, sheep grazing has been found to limit CO2 stimulation of grassland productivity by selectively consuming legumes and forbs, plants with the greatest growth responses to CO2 (Newton et al., 2014). The issue of uncertainty is even more significant for grass yield projections than for crops, as reference data are less available for models’ development and evaluation.

2.1.1.4. Impacts of pests, pathogens, weeds and pollinators on yields

The effect of climate hazards on pests (e.g. insect pests, pathogens), weed outbreaks and pollinators can have significant consequences for animal feed availability as reviewed in Myers et al. (2017).

Pests, pathogens and weeds are estimated to currently reduce the production of major crops by 25–40% (Flood, 2010). Increases in temperature increase winter survival of insect pests and rates of herbivory (Bale et al., 2002), and alter the spatial distribution of pests and pathogens. Bebber et al. (2013) reported an average poleward shift of pest and pathogen distribution of 2.7km per year since 1960, though there is substantial variation among taxonomic groups. Under climate change, the spatial mismatches between pests and natural predators may be exacerbated in some regions, weakening biological control systems (Selvaraj et al., 2013). In some instances, weather extremes can weaken crop defences and create niches for pests and weed outbreaks (Rosenzweig et al., 2001). In other cases, extreme events can reduce pests and weeds, and as such support crop establishment and growth (Young, 2015). Recent intense desert locust outbreaks across East Africa, Asia and the Middle East have been linked to a series of cyclones causing warm and wet conditions (Salih et al., 2020). In Ethiopia, as per April 2020, nearly 200,000ha of cropland were damaged by the insects, leading to the loss of over 356,000 tons of grain and thousands of tons of crop residues, a key livestock feed in the country (FAO, 2020). An additional 1.3 million hectares of pasture were affected, reducing pastoral areas by as much as 61% in the Somali region. Increases in eCO2 concentrations may also influence the weed composition and crop defences in complex ways (Zvereva and Kozlov, 2006), including reducing the effectiveness of herbicides (Ziska and Goins, 2006; Ziska and George, 2004). Shifting pest and disease patterns may increase the use of pesticides, some of which (i.e. dioxins) can pass on to animal products. These toxins can remain in soils for extended periods, and can contaminate animal feeds and water sources, particularly in conditions with alternating periods of drought and floods that are more likely with climate change (van der Spiegel et al., 2012).

More unstable weather, including more humid and cloudier conditions will lead to more on-farm post-harvest losses of animal feed, especially in developing countries with hot climates where most smallholders rely on the sun to dry their crops and forages before storage (Hodges et al., 2011). Contamination by toxins will likely also be higher (van der Spiegel et al., 2012). For instance, maize and sorghum can become contaminated by aflatoxins, particularly in drought conditions. While concentrations at harvest are usually not poisonous, the storage of grains under damp or poorly aerated conditions can lead to mould and the poisoning of animals and consumers. Increases in pest infestation frequency or intensity under climate change will also result in higher crop losses where storage facilities are inadequate.

While all grasses and most staple food grains such as maize, wheat, rice and sorghum are wind or self-pollinated, some crops used as livestock feed in industrial and mixed crop-livestock systems are animal pollinator-dependent to varying levels (e.g. soybean, cowpeas, pigeon peas, broad beans, rapeseed, oilseed rape, oil palm and some vegetable and fruit crops) (Klein et al., 2007). Climate change impacts on pollinators include changes in the abundance and distribution of both flowering plants and pollinators (Abrol, 2011; Hegland et al., 2009; IPBES, 2016; Memmott et al., 2007), and the timing of flowering and pollinator emergence and migration (Parmesan and Yohe, 2003), causing a mismatch in pollinator availability and crops to be pollinated. This contributes to reductions in the breadth and nutritional value of feed for pollinators (Ziska et al., 2016), which in turn decreases pollinator abundance. Increases in eCO2 concentrations also affect the nutritional value of key forages for pollinators (Ziska et al., 2016). While the net effect of climate change on pollinators remains uncertain, studies indicate that a reduction in animal pollination would decrease yields of numerous pollinator-dependent food crops (Klein et al., 2007).

2.1.2. Water availability and quality

Water is used at various stages of the livestock supply chain: to grow feed; for animal consumption and cooling; to produce electricity, fertilizers, pesticides and fuel; and to clean animals and infrastructure (see detailed inventory in FAO (2019b)). Most water in the livestock value chain is used for feed production, accounting for over 90% of water consumption in many systems (Legesse et al., 2017; Mekonnen and Hoekstra, 2012). The amount of water required for livestock consumption varies with local climate conditions, with higher consumption under hot conditions (Ward and McKague, 2019). For instance, once air temperatures exceed 30°C, the expected poultry drinking water intake can increase by 50% above normal rates. In arid and semi-arid regions, the increased frequency, intensity and duration of droughts are a significant challenge for sustaining water supply to livestock and feed production.

Climate change is also projected to reduce raw water quality (Jiménez Cisneros et al., 2014), which can decrease animal water intake, feed intake and health (The State of Victoria, 2018; Valente-Campos et al., 2019). Poor local water quality can be caused by warmer temperatures, sea level rise, or higher sediment, nutrient and pollutant loadings due to heavy rainfall. Reduced dilution of pollutants during droughts and disruptions of treatment facilities during floods are also a concern. High salt concentrations in water and feed depress livestock feed intake and production (Masters et al., 2006; Sharma et al., 2016). The tolerance of animals to different levels of total dissolved solids (including inorganic salts) in water varies considerably among varied species with poultry, buffaloes and dairy cattle having lower tolerance levels (desirable maximum concentration for healthy growth of 2000–2500ppm) than beef cattle, sheep and pigs (4000ppm, Sharma et al., 2016).

2.1.3. Shifts in agricultural lands

The effects of climate change on resources, as mentioned in this section, will continue to lead to shifts in the global agricultural area as well as changes in seasonality and crop and livestock suitability. For example, King et al. (2018) estimate a northward shift of the feasible agriculture zone by up to 1200km by the end of the century, although many of these areas are associated with highly variable water balances. The world's drylands have expanded over the last 60 years, mostly in semiarid regions, and are projected to expand further during this century (Huang et al., 2017). In sub-Saharan Africa, more than 20% of the mixed crop-livestock system in arid and semi-arid regions is projected to become unviable for crop agriculture by mid-century (Jones and Thornton, 2009). Grass yields may overall be less impacted by climate change than crop yields, which may favour grazing systems and potentially change the current trend towards more intensive systems (Havlík et al., 2015).

2.2. Animal health and production

The impacts of climate change on animal growth, production and welfare include impacts mediated by reduced feed intake, direct physiological and metabolic effects, and changes in behaviour (e.g. Filipe et al. (2020), Gauly and Ammer (2020), Lara and Rostagno (2013), Lees et al. (2019), Nardone et al. (2010), Polsky and von Keyserlingk (2017), Romo-Barron et al. (2019), Saeed et al. (2019), Sevi and Caroprese (2012), Tedeschi and Fox (2020)). The potential impacts of climate change on livestock health and production are already well-reviewed in the literature, and we highlight key effects below.

2.2.1. Heat stress

Heat stress is primarily caused by exposure to high ambient temperatures and relative humidity, which limit the capacity of livestock to shed heat to their environment, on farm or during transportation (see section 2.3.1). Vulnerability to heat stress varies according to species, breed, life stage, genetic potential, nutritional status, size, level of insulation (hide thickness or distribution of feathers) and previous exposure of the animal, with animals with high energy demands (i.e. high-yielding individuals and breeds) most susceptible (Bernabucci et al., 2010; Rashamol et al., 2019; Saeed et al., 2019). For example, dairy cows are more susceptible than beef cattle, and temperate Bos taurus breeds are more susceptible than tropically adapted Bos indicus cattle and their crosses (Polsky and von Keyserlingk, 2017).

In most cases, the impact of heat stress is reduced productivity and animal welfare. However, under severe or prolonged conditions, mortalities will also occur. Reduced feed intake is one of the first and biggest consequences of heat stress, leading to declines in growth rates and production of milk or eggs.

Heat stress also affects livestock production through changes to fertility and susceptibility to disease. In both mammals and poultry, the impacts on fertility are caused by reduced ovarian function, reduced motility of spermatozoa, and inhibition of embryonic development (Nawab et al., 2018; Polsky and von Keyserlingk, 2017; St-Pierre et al., 2003). Cattle also show reduced expression of estrus behaviour, further reducing the chances of reproductive success. In addition, heat stress may depress immune function and the effectiveness of some vaccines (Bagath et al., 2019; Hirakawa et al., 2020), increasing the incidence of livestock disease. The reverse is also true, with animals showing clinical signs of disease more susceptible to heat stress (Gaughan et al., 2008). While covered to a limited extent in the literature, there is some evidence that high temperatures may decrease the ability of mammalian herbivores to detoxify plant secondary compounds (Dearing, 2013; Moore et al., 2015). While most work to date has been done with rodents and wildlife (e.g. Kurnath et al. (2016)), high environmental temperatures have been shown to increase the susceptibility of cattle to poisoning from ergot alkaloids in tall fescue (Aldrich et al., 1993).

The impacts of heat stress on livestock can be both immediate and long-lasting, and can also affect offspring exposed to heat stress in utero. Research in dairy cattle and pigs has shown that heat stress in utero reduces milk yield at first lactation (Dahl et al., 2016; Monteiro et al., 2016), and alters nutrient partitioning and carcass composition (Boddicker et al., 2014; Johnson et al., 2015). Animals exposed to heat stress in utero may also be better adapted to heat stress conditions at maturity, with Ahmed et al. (2017) reporting that cows exposed to heat stress in utero are better able to regulate core body temperature.

Heat stress can also impact the quality of animal products, reducing the size of eggs and thickness of eggshells (Mashaly et al., 2004), decreasing the fat and protein content of milk (Bernabucci et al., 2002; Sevi and Caroprese, 2012), and changing the colour and water-holding capacity of both red and white meat (Gonzalez-Rivas et al., 2020). These changes can make livestock products less appealing to customers, increase wastage, and reduce the price that producers receive for their products. Selection of breeds that are better adapted to high temperatures can also have implications for product quality; while Bos indicus cattle are better adapted to high temperatures and humidity than Bos taurus, their meat is less tender (Crouse et al., 1989; Johnson et al., 1990), and tends to score lower in meat quality assurance schemes such as the Meat Standards Australia index, receiving lower prices in some markets.

Heat stress, and other stressors such as feed withdrawal can also cause the dissemination of enteric pathogens such as Salmonella, Escherichia coli O157:H7, and Campylobacter from livestock into human food, and as such is also a major health concern. Indeed, these external stressors can increase animal pathogen carriage and shedding as reviewed in Gonzalez-Rivas et al. (2020). The underlying mechanisms have however not yet been fully explained.

2.2.2. Other impacts

In addition to heat stress, the increased frequency and intensity of storms, fires, floods and cold waves can have significant repercussions for the livestock sector. Extreme flooding followed by cold weather in northern Australia in early 2019 caused the loss of approximately half a million livestock (mostly cattle, but also sheep, goats and horses) (Queensland Government, 2019). Similarly, in October 2018, Hurricane Michael destroyed an estimated 84 chicken houses and killed over 2 million chickens in Georgia, U.S (Georgia Department of Agriculture, 2018). In geographic areas with cold winters, such as in the Northeast US, warmer temperatures may also reduce animal cold stress and maintenance energy requirements, as well as housing heating (Hristov et al., 2018; Toghiani et al., 2020).

Climate change may also impact infectious livestock diseases by changing their spatial distributions, affecting annual and seasonal cycles, altering disease incidence and severity, and modifying susceptibility of livestock to illness (Bagath et al., 2019; Filipe et al., 2020; McIntyre et al., 2017; Patterson and Guerin, 2013; White et al., 2003). Many infectious pathogens that cause disease in livestock are sensitive to changes in climate, primarily moisture, rainfall, temperature and particulate matter – many of these diseases are zoonotic (McIntyre et al., 2017). Some pathogen transmission routes are also more climate-sensitive than others. Vector-borne, foodborne, water-borne and soil-borne pathogens are the most likely to be affected by climate change, while those transmitted directly or by fomite are the least likely to be affected by climate change (McIntyre et al., 2017). In particular, the quantity and spread of insect vectors such as flies, ticks and mosquitoes, but also wild birds, rodents and mammals which can transmit disease to farmed poultry, pigs and ruminants is of concern to the livestock sector.

It is also probable that increases in tropospheric O3 will affect animal health (Menzel, 1984), though increases in mortality are likely smaller than those reported for heat stress events (Egberts et al., 2019).

2.3. Processing operations, storage, transport, retailing

2.3.1. Live animal transportation

Livestock are often transported long distances by road or sea to markets and slaughter. Climate change, particularly increases in temperature and climate-driven disruptions in the transport network, may increase the risk of heat stress during transport, which in turn, can contribute to poor animal welfare and deaths (Caulfield et al., 2014; Collins et al., 2018). This aspect is not often considered in heat stress literature reviews. High stocking density of animals during transport increases ambient temperature and humidity, and reduces ventilation, decreasing opportunities to shed heat loads. In addition, animals transported by truck usually do not have access to water on-board and can become dehydrated faster during hot conditions. In response to this, the live export of sheep from Western Australia to the northern hemisphere is currently prohibited in summer (June to September) (Australian Department of Agriculture and Water Resources, 2019). Increased scrutiny of animal welfare in other regions may result in additional restrictions on animal transport such as when animals can be moved (time of day and year), and the duration of transport.

2.3.2. Built-up capital

Higher temperatures, increased humidity, increased frequency of extreme weather events, and rising sea levels put additional stress on built-up capital (machinery, transportation infrastructure, electricity networks, telecommunications, etc.). This will lower productivity across many economic sectors, increase construction, operating and maintenance costs, and may shorten the lifespan of critical industry infrastructures (Lloyds, 2015; Schweikert et al, 2014a, 2014b; Wang et al., 2012).

Damage and degradation of key transportation infrastructure such as roads, railways, and port infrastructure are of particular concern for the livestock sector (Markolf et al., 2019). For example, in Australia, 95% of livestock are transported domestically by road from farms to central points such as saleyards, feedlots, abattoirs or ports of embarkation for live export (Fisher et al., 2006). In early 2019, cattle producers in northern Australia reported over 29,000km of farm roads destroyed by extreme flooding, in addition to the loss of tools, machinery, homes and 22,000km of fencing (Queensland Government, 2019). At a sea level rise of 1.1m (high end scenario for 2100), the Australian Government projects that coastal assets at risk from the combined impact of inundation and shoreline recession are greater than AU$226 billion and include between 27,000 and 35,000km of roads and rail, with a value of between AU$51 and AU$67 billion (Commonwealth of Australia, 2011). In the EU, almost 90% of external trade is transported by sea (Suárez-Alemán, 2016) and studies found that 64% of all European seaports are expected to be affected by inundation events by 2060, caused by global mean sea level increases and combined effects of tides, local waves, and storm surges (Christodouloum and Demirel, 2017). Reduced polar ice may offer new trade routes that could reduce travel time between Europe and the Northern Pacific (Lindstad et al., 2016), and could reduce the reliance on several key trade passages (e.g. Panama Canal, Suez Canal, and Strait of Malacca) but could also threaten new marine ecosystems through the introduction of invasive species (Miller and Ruiz, 2014).

2.3.3. High temperatures, product distribution and storage

Increasing temperature results in worsening conditions for storage, which leads to degradation of food quality and shelf-life, increased wastage and increases the likelihood of the proliferation of microbes and fungi (i.e. E. coli, salmonella, aflatoxin and mycotoxin producing organisms, etc.), especially under humid conditions (van der Spiegel et al., 2012). Even in developed economies, increased temperatures have been found to increase the likelihood of food poisoning, suggesting that increasing temperatures will lead to reduced food quality and safety (D'Souza et al., 2004). For instance, chilled (below 10°C) storage life is halved for each 2–3°C rise in temperature (James and James, 2010). In addition, where climate change results in higher levels of microorganisms on meats and produce prior to processing, storage temperatures may need to be lower to preserve required shelf-lives.

Climate control (air conditioning and refrigeration) can mitigate some of the negative impacts of climate change on food processing and distribution. However, this leads to increased energy usage and costs, and current levels of refrigeration are already estimated to contribute 1% of global CO2 emissions (James and James, 2010). Sarhadian (2004) found that an increased ambient temperature from 17 to 25°C resulted in an 11% increase in average power consumed in a catering establishment. As a food moves along the cold chain it becomes increasingly difficult to control and maintain its temperature. This is because the temperatures of bulk packs of refrigerated product in large storerooms are far less sensitive to small heat inputs than single consumer packs in open display cases or in a domestic refrigerator or freezer (James and James, 2010). Fresh milk, in particular, requires significant amounts of cooling and could be the most impacted by increasing costs, or deterioration of cold chains.

Globally, the use of refrigeration is also far below what is optimal, with only 10–20% of perishable value chains refrigerated (Coulomb, 2008). Low- and middle-income (LMIC) countries in already hot environments are particularly vulnerable, due to limited cold chain development. These are also the regions that have seen the largest increases in consumption of livestock products in recent years (FAOSTAT, 2020; Herrero et al., 2018). It is estimated that if LMIC acquired the same level of refrigerated equipment as that in high income countries, wastage of perishable food would decrease by over 200 million tonnes, or 14% of the current consumption in these countries (International Institute of Refrigeration, 2009).

2.3.4. Extreme events, products distribution and storage

Changes to the frequency, intensity and duration of extreme events will impact supply chains by increasing uncertainty in input flow and damaging infrastructure. This will make trade patterns less regular, increasing reliance on complex logistic systems. Extended droughts reduce hydropower generation, and can lead to increased wildfires, both of which can lead to brownouts, a particular threat to supply chains heavily reliant on temperature controls to maintain food quality and safety. Extreme events can also damage key transportation (road, rail, ports) infrastructure, limiting the distribution of products. Both low and high flow can affect the navigability of key inland waterways like the Mississippi River (Olsen et al., 2005) and Great Lakes in North America (Attavanich et al., 2013). For instance, a major drought in 1988 drastically reduced barge movement along the Mississippi a critical corridor for transporting grains in the USA to the port of New Orleans (Changnon, 1989).

In many countries the supply chain for livestock products is highly concentrated and coordinated, and disruptions anywhere along the supply chain can have impacts throughout the chain. For example, the COVID-19 outbreak in the USA prevented the processing of pigs, which forced pork farmers to either hold pigs, or cull them to avoid paying to feed them (BBC, 2020). There is more scope for keeping animal stocks for beef, but any culling in herds also has a longer impact on animal stocks due to lower reproductive capacity of cattle compared with chickens or pigs.

2.4. Livestock product consumption

We highlight in previous sections the potential impacts of climate change on the supply and processing of animal products, as well as on the quality and safety of animal products through contamination with pathogens or pesticide (sections 2.1.1.4, 2.3.3) and reduced nutritional quality and sensory appeal (section 2.2.1).

The social licence of the livestock sector is increasingly challenged by the rising awareness of climate change and livestock impacts. These societal changes are also influenced by animal welfare and environmental concerns, and increasingly supported by civil society, governments and other institutions. Changing social norms are likely to impact diets, especially in high-income countries (Godfray et al., 2018). Research also found that temperature changes can affect dietary preferences of consumers (Motoki et al., 2018; Needs et al., 1993), although the effects on the consumption of livestock products remain largely understudied. Changes in the price of animal products and relative magnitude of price changes across commodities will also alter consumption patterns (Muhammad et al., 2017; Valin et al., 2014). These changes will be driven by product availability and quality, as well as costs along the supply chain. The rate of growth in gross domestic product, which can be influenced by climate change (Burke et al., 2015), will also alter consumers' purchasing patterns. Indeed, food price sensitivity tends to fall as incomes rise, and a smaller share of income is spent on food (i.e. Engel's Law; Clements and Si (2018)). Sensitivity to primary commodity prices may also decline as supply chains complexify: increasingly the cost of the food is driven by value-added processing, packaging, and branding. For example, the farm share of the food value added in the European Union has fallen from 31% in 1995 to 24% in 2005 (European Commission, 2009).

2.5. Labour

Labour availability and productivity are key determinants of the efficiency of the livestock supply chain. This is especially the case in less developed regions that do not rely on mechanization. However, human performance and health can be limited by a broad range of direct and indirect effects of climate change as reviewed in previous studies (Patz et al., 2005; Smith et al., 2014). These impacts relate to temperature, floods and storms, ultraviolet radiation, infections, air quality, nutrition, occupational health, mental health, violence and conflicts. For instance, in 2010, inhalation of climate-altering pollutants other than CO2 caused more than 7% of the global burden of disease (Smith et al., 2014). We further detail below two key drivers of labour productivity of particular concern to agricultural labour: heat stress and diseases.

Above 24–26°C, labour productivity declines (International Labour Organization, 2019). At 33–34°C, a worker operating at moderate work intensity loses 50% of their work capacity. In addition, exposure to excessive heat levels can lead to heatstroke and sometimes fatal outcomes. Productivity may decrease by 11–27% by 2080 in hot regions such as Asia and the Caribbean (Kjellstrom et al., 2009), and labour productivity for high intensity work declines by up to 31%–⁠38% (RCP 4.5–⁠RCP 8.5) in Southeast Asia and the Middle East by 2050, relative to a 2050 baseline without climate change (Knittel et al., 2020). In some areas, 30–⁠40% of annual daylight hours will become too hot for work to be carried out (Kjellstrom et al., 2016). Literature focussed on agricultural labour emphasises the challenges caused by physical labour under high temperatures, with dehydration often being the main cause of lost labour productivity (Wagoner et al., 2020; Wästerlund, 2018). Heat stress is of particular concern in production systems that are dependent on high inputs of human labour and located in environments that are already hot, such as smallholders in sub-Saharan Africa, as highlighted in Frimpong et al. (2017) and Yengoh and Ardö (2020). Foreign workers, who represent a large part of the agricultural labour force in some countries such as in the U.S. are also particularly at risk. Foreign farmworkers may have entered the country illegally, often do not speak the local language, have low education levels and incomes and limited access to health care. They often feel they have little control of their workplace, and tend to be reluctant to complain about unsafe work environments (Lambar and Thomas, 2019).

In addition to heat stress, communicable diseases such as Malaria, dengue fever, tick borne encephalitis, borreliosis, salmonellosis, cholera and bluetongue are all affected by climate change and affect the health of people working along the livestock supply chain (Caminade et al., 2019; Wu et al., 2016; World Health Organization, 2014). For instance, excessive rainfall and high temperature results in an increased transmission rate, reproduction rate and the proliferation of the Plasmodium species that causes malaria (Braide et al., 2020), and outbreaks of leptospirosis in Sub Saharan Africa (Lau et al., 2010). In areas where there is inadequate access to water and sanitation, flooding increases people's exposure to salmonellosis and cholera. Conversely, a drying climate and extended periods of drought in Australia have been conducive to the spread of Q-fever, with the disease-causing bacteria Coxiella burnetii able to survive and travel long distances on dust particles contaminated with animal faeces or birth fluids (Archibald, 2019). Warming temperatures in northern latitudes have also been reported to favour the distribution of anthrax, particularly in association with increased livestock densities (Walsh et al., 2018). Populations in close contact with livestock and wildlife are particularly at risk of zoonotic pathogens, with climate change further contributing to the risk of zoonosis (McIntyre et al., 2017).

2.6. Prices

Climate-related challenges from resource availability to consumption levels mentioned above can result in rises in costs of water, feeding, housing, storing, transport, retailing and insurance, which negatively impact actors throughout the livestock supply chain. An ensemble of crop and economic models with various representations of the global food system and associated assumptions (e.g. related to technological change, land-use policies or consumption patterns) projects a 1–29% increase in cereal prices by 2050 across Shared Socio-economic Pathways 1, 2, and 3 under climate change (RCP 6) (Hasegawa et al., 2018). This would impact consumers globally through higher food prices, with effects varying regionally. The price of animal products is also projected to increase, but projected prices are about half that of cereals, highlighting that in these models, the impacts of climate change on animal products are felt mainly through changes in costs and availability of feed. These results also highlight scope for feed substitution within the livestock sector. Most climate studies have focused on price levels and not on price volatility. However, with increased spatial and temporal variability in production and supply chain efficiency, it is probable that food prices will be more volatile under climate change (Mbow et al., 2019). The international agricultural trade system is dominated by regional trade (i.e. within Europe and Mediterranean Basin, North America, etc.), with trans-regional trade dominated by a few major international exporters (e.g. soybeans from Brazil, maize from the US, rice from Southeast Asia, lamb from New Zealand-lamb, etc., Herrero et al., 2018). Major shocks to these major producers could dramatically increase the volatility of international markets and reduce the reliable supply not only of food but also of feeds, which are traded at much higher volumes than livestock products. Furthermore, it can drive policy responses to protect domestic supply that can exacerbate food crises (Headey and Fan, 2010).

2.7. Social dynamics and systems change

Climate change is contributing to a range of societal changes with implications for the livestock sector. As noted above, changes in climate conditions will challenge productivity, economic viability, structure, and attractiveness of livestock-related enterprises with consequences on the availability, quality, affordability, safety and sensory appeal of livestock products. Climate change will likely amplify existing inequalities with particularly negative impacts on small-scale enterprises, which may struggle to remain competitive (Hallegatte and Rozenberg, 2017; Islam and Winkel, 2017). Social learning, the process of learning from and of responding to experience, will also shape social-ecological systems (Morrison et al., 2017). The mitigation and adaptation actions that food system actors themselves take may also affect social dynamics. For example, rising awareness of the negative consequences of climate change and the sector's environmental footprint may contribute to shifting social norms, particularly in high-income countries, that may favour less resource intensive production (Godfray et al., 2018), and encourage the further development of novel protein products (plant-based meats, cultured meats, etc.). All these changes will impact both local and global food systems dynamics and food environments.

3. Vulnerability of livestock-based socio-economic systems

The ability of the livestock sector to cope with or adapt to climate change, i.e., its adaptive capacity, depends on a range of socio-economic, political, institutional and environmental factors from local to global scales that interact in complex ways. Well known demographic, social and economic factors are driving change in the livestock sector. These factors are briefly presented below and treated in more depth in the Supplementary Information.

A growing human population, together with increasing incomes and shifts in dietary preferences, predominantly in low- and middle-income countries, are driving continued growth in demand for livestock products (especially poultry and pork) and resources needed to produce them, including land and water (FAOSTAT, 2020; Herrero et al., 2018). Potential enforced production systems transitions due to changes in land suitability will affect the food production landscape (Thornton et al., 2019). Large-scale demand for plant-based diets and production of synthetic livestock products may also alter livestock production systems dynamics (Springmann et al., 2018). The development of such food alternatives will be influenced by changing social norms and opportunity-costs of different food production systems under climate change and other global trends. In addition, competition between crop as livestock feed and crop for direct human consumption or biofuel production are increasing (Muscat et al., 2020). Human migration, partly induced by climate change, can also increase local pressures on scarce natural resources (FAO, 2018b). Rising food safety standards and governments' investments in infrastructure, price support schemes, taxes, credits, subsidies, input and output quotas and the health system, will also influence the vulnerability of livestock supply chains. For instance, the lack of infrastructure and access to knowledge, land, financial services, markets, new dynamic sectors and policy dialogues in rural areas is making the agricultural sector unattractive to youth, with negative implications for the agricultural sector's performance and adaptive capacity (FAO, 2014; Proctor and Lucchesi, 2012; White, 2012). Production challenges associated with harsh climates can also increase youth reluctance to enter or remain in agriculture. Discrimination based on gender, ethnicity, caste, and wealth impedes participation in markets, legal recognition of land and asset ownership, and other rights that play key roles in the ability of stakeholders to adapt to climate change. Civil conflict and institutional disregard of traditional knowledge, institutions and customary practices can also weaken the resilience of livestock systems (FAO, 2018b; IPCC, 2014b; Nielsen et al., 2020). While some of the political and institutional forces may contribute to persistent trends (e.g. continued exclusion of pastoralists from political processes), others may result in dramatic shifts in the near future and greatly alter the ‘vulnerability landscape’. The pace at which governance factors can change has been exemplified by governments' actions to stop the spread of diseases (e.g. H5N1 or COVID-19).

People in different production systems have differentiated exposures and vulnerabilities. Table 2, Table 3 highlight the exposure of people in different regions and production systems based on economic and livelihood indicators of the livestock sector. We summarise in Table 4 key vulnerabilities and adaptative capacity characteristics of people in grazing-only systems, mixed crop-livestock systems and industrial systems, as defined by Seré and Steinfeld (1996) and mapped in Robinson et al. (2011). These vulnerabilities are also discussed in Rivera-Ferre et al. (2016), and a focus on adaptation in mixed systems is provided in Thornton and Herrero (2015, 2014).

Table 2

Land-based livestock economic and livelihood indicators – Total protein production and value of production per region and globally, and (shaded) relative contribution from different livestock production systems. The darker the shade, the higher the contribution. Livestock production systems as defined by Robinson et al. (2011). Calculations presented in the Supplementary information.

Impacts of climate change on the livestock food supply chain; a review of the evidence (5)

Table 3

Land-based livestock economic and livelihood indicators at regional level. Calculations presented in the Supplementary Information.

Impacts of climate change on the livestock food supply chain; a review of the evidence (6)

Table 4

Some current key vulnerabilities and strengths of people in different livestock production systems. Livestock production systems as defined by Seré and Steinfeld (1996) and mapped in Robinson et al. (2011): livestock grazing systems (LG) and mixed systems (M) in arid (A), humid (H) and temperate (T) regions. Table adapted from Rivera-Ferre et al. (2016).

Grazing systemsMixed crop-livestock systemIndustrial system
Vulnerability• Political marginalization
• Land encroachment
• Land degradation
• Land fragmentation
• Remoteness
• Reliance on physical labour related to limited mechanization
• Lack of financial capital and alternative economic options
• Conflicts (civil conflicts, conflicts over resources)
• Land ownership and tenure arrangements (e.g. communal land tenure can limit land and infrastructure improvements)
• Limited mobility
• Land degradation
• Land scarcity especially from urban expansion
• Rising food safety standards
• Population growth
• Economic margins often small and financial capital often low, resulting in lock-in
• Economic competition favouring cropping
• Co-managing price and climate variability
• Learning and capital demands from having multiple farm components
• Labour supply for peak periods of activity
• Shrinking farm sizes
• Dependence on external inputs and hired labour
• Energy intensive
• Difficulties in re-locating built-up capital
• Narrow gene pools in livestock and input crops
• Large, high-yielding animals are more susceptible to heat stress and disease
• High-yielding crops are often more sensitive to heat and water stress
• Challenges in waste disposal and animal welfare impacting on social licence to operate
• Susceptibility to disease outbreaks
• Low economic margins
• Operating close to or at maximum physiological and financial limits
• Integrated in highly efficient value chains
Adaptation capacity• Mobility to adapt to spatial and temporal climate variability
• Family labour
• Communal land and social collaboration
• Local knowledge of diverse resources
• Capacity to add value to marginal land via provision of ecosystem services
• Wide livestock gene pool
• Recycling plant nutrients
• Transformation to mixed systems
• Off farm income
• Integration of agriculture and livestock
• Capacity to use crop residues
• Often private land, hence have agency
• Flexibility in crop-livestock allocation and other decisions
• Diversification
• Family labour
• Wide livestock and forage gene pool
• Recycling plant nutrients
• Flexibility in allocating produce to subsistence or market
• Off farm income
• Access to global feed and input supply chains
• Access to credit and modern technology
• Access to global consumer market
• Capital mobility and exploiting economies of scale
• Control of many aspects of the system
• Good information systems (climate, financial, supply) allowing rapid responses

Some characteristics of grazing systems compared with other food production systems are their usual remoteness from population centers and limited access to information, markets, capital, labour and veterinary services. These characteristics, together with market price fluctuations and social considerations (e.g. benefit perception), limit opportunities for management options that could mitigate the effects of climate change and high climate variability such as timely stock adjustments. People in these systems, often both geographically and politically distant from policy makers, tend to be marginalised and receive limited investments from governments and businesses (Godber and Wall, 2014; Marshall, 2015; Nielsen et al., 2020; Sayre et al., 2013; Thomas and Twyman, 2005). Grazing systems are often located on land that is poorly suited for cropping, notably due to high climate variability. Herd mobility and collective resource management have been key strategies to take advantage of the spatial and temporal variation in water and forage availability. However, changes in land tenure and increasing landscape fragmentation have become a major concern for grazing systems resilience in many parts of the world (Hobbs et al., 2008; Reid et al., 2014). For instance, in Inner Mongolia, the culture had adapted to the harsh climate by using mobility, cooperation, and reciprocity strategies as described in Dalintai et al. (2012). However, collectivization between the 1950s and mid-1980s followed by market reforms in the early 1980s disrupted these traditions and weakened herd mobility. The state-driven nomad sedentarisation projects in China (Hruska et al., 2017) and transitions from communal to semi-commercial land tenures in southern African rangelands (Dube and Pickup, 2001) have also limited adaptive capacity of people in these regions. In Australia, where most grazing land is owned or leased, it is common practice to move cattle from a droughted area to other privately run properties that have adequate pasture, and graze for a fee (“agistment” strategy; McAllister, 2012; McAllister et al., 2006). The shortage of productive land during widespread droughts can however greatly limit the possibility and economic viability of such a mobility strategy.

Farmers in mixed crop-livestock systems often have more limited economic opportunities as compared to cropping systems (Rivera-Ferre et al., 2016). This is particularly the case for small-scale systems. Shrinking farm sizes limit opportunities for management adaptations and more stringent food safety standards limit the ability of smallholder farmers to enter expanding markets as these standards create new knowledge requirements, additional investments and stronger linkages between producers and buyers (Humphrey, 2017). The literature has extensively covered the impacts of climate change on crop enterprises in mixed systems, and to a lesser extent, the impacts on livestock enterprises (Thornton and Herrero, 2015), although the focus has often been on yields rather than livelihood outcomes or other sustainability dimensions (Ricciardi et al., 2020). However, relatively little is known about how crop-livestock interactions may be affected or offer buffering capacity to help smallholders adapt to climate change (Thornton and Herrero, 2015).

Industrial systems are generally based on high-yielding animals, which are more susceptible to heat stress, so require greater investment in infrastructure to insulate them from climate extremes. Compared to more extensive systems, the close physical proximity of animals to each other and humans can increase the possibility of disease outbreaks. These outbreaks can however be more quickly identified and contained. The limited contact with wildlife also reduces the risk of diseases from endemic carriers. Industrial systems tend to be energy and capital intensive which can make them vulnerable to disruptions to energy supply. Due to greater integration in efficient value chains, they are also more vulnerable to disruptions to transportation and demand given more limited capacity for long run inventory.

Overall, livestock ownership itself can support farmers resilience in times of climate change and climate extremes. This is particularly the case in smallholder farming systems, where livestock plays multiple roles beyond producing food for the market (Alary et al., 2015; CIRAD, 2016). These roles include provision of wool, hides, skins, manure, animal traction, transportation and food for home consumption. Livestock can also contribute to the diversification of income and accumulation of capital savings, and as such can also be used as a risk reduction strategy by vulnerable farming communities. In particular, livestock production is often more resilient to high climate variability or short length of growing periods than crop production (Robinson et al., 2011; Sloat et al., 2018). Animals can also be sold in times of climate shock to help compensate for crop failures, loss of income or additional expenses (WFP, 2019). Diversified livestock systems, including the ownership of multiple animal species and breeds, improve dietary adequacy and diversity (Jones, 2015; Murendo et al., 2018; Ritzema et al., 2019; Romeo et al., 2016) and food security (Bahadur KC et al., 2016). Shifts in animals breeds can serve as a climate adaptation strategy (e.g. shifts from cattle to camels in East and North Africa or from British Bos Taurus to Zebu Bos indicus in Australia, Bortolussi et al., 2005; Faye et al., 2012; Kagunyu and Wanjohi, 2014; Watson et al., 2016). Nevertheless, climate change may challenge the viability of livestock systems in some contexts, given environmental, socio-economic and institutional constraints to adaptation.

4. Societal risk of impacts on livestock supply chains - at the intersection of hazards, exposure and vulnerability

The risks of climate-related impacts arise from the interaction of climate-related hazards (including hazardous events and trends) with exposure of the livestock supply chain to these hazards and the vulnerability of the socio-economic systems within which they are embedded (Fig. 1).

Worldwide, rangeland communities that are projected to see the most negative impacts of climate change on vegetation are also amongst the communities most vulnerable, according to a range of socio-economic vulnerability indicators at local and national levels (see Supplementary Information, Godde et al. (2020)). Across a range of dimensions, including food security, health, socioeconomics, and governance, Godber and Wall (2014) found that the region with livestock-based food production at most risk was South Asia. They found that LMICs were generally more at risk than higher income regions, with the top decile of nations at risk located in Sub-Saharan Africa, seven of which were in Eastern Africa. The potential impacts emerging from such risks will affect all dimensions of sustainable livelihoods⁠—human, social, natural, physical, and financial capitals (Fig. 2).

Climate-related risks will be context specific, but also greatly influenced by global socio-economic trends and shocks. Differences in risks arise from climatic and non-climatic factors and from multidimensional inequalities, often due to uneven development processes. These context specificities highlight the importance of renewed attention to diversity within the livestock sector and its multiple socio-economic and environmental contributions. For instance, in Oceania, livestock represents 70% of agricultural added-value (Table 3) and over 90% of livestock feed is from grazing (Table 2, Table 3). Most of Oceania's grazing lands (mainly in Australia) are subject to high and increasing climate variability and frequent and intense fires, droughts and floods (Nielsen et al., 2020). As such, the viability of the sector will greatly depend on the success of the strategies implemented to cope with high climate variability. These strategies will need to account for geographic remoteness, limited economic margins, and over-grazing. Increasing climate variability and drying trends on grasslands is also a key concern in arid and semi-arid African regions (where 60% of feed is from grazing), as are changes in the length and start of the crop growing seasons (over 70% of protein production comes from ruminant mixed systems and grain-fed chicken; Table 2, Table 3; Thornton et al., 2011). Socio-economic and institutional support, including social safety nets, is urgently needed considering the multiple roles livestock plays in such regions. Overall, small-scale farmers in environments that are already hot and with limited resources for adaptation will be the most at risk. We estimate that 228.6 million people worldwide live in low-income countries in grazing-only and mixed crop-livestock regions in arid environments. Agriculture represents approximatively 57% of total employment in these countries (see Supplementary Information for calculations). Although the share is likely to decline in the future, in absolute terms, the number of vulnerable people employed in agriculture will remain high.

The impossibility of predicting complex interconnected biophysical and social systems far into the future challenges the evaluation and prioritisation of climate risk mitigation and management strategies. Given this fundamental uncertainty, it is crucial to consider a wide range of plausible futures and climate outcomes, which while it may not allow us to predict the future, can allow us to assess the range of possible impacts and to test the robustness of policies and interventions across alternative contexts (IPCC, 2014c).

5. Conclusion

The rapid growth of the livestock sector and its various contributions to the economy and human livelihoods highlight the importance of better understanding the impacts of climate change on livestock.

While it is certain that climate change will impact the sector throughout the food supply chain⁠—from farm production to processing operations, storage, transport, retailing and human consumption⁠—large uncertainties remain as to the nature, extent and magnitude of these impacts. These uncertainties relate to the future climate as well as exposures and responses of the interlinked human and natural systems to climatic changes over time. For instance, the impact of climate change on agricultural labour, on the livestock supply chain from farmgate to retailers or on livestock products prices and consumption behaviours have received limited attention in the past. The impacts of climate change on livestock feed production and quality, especially forages, have also been understudied as compared to key staple grains. In addition, the rates of change in global socio-economic and environmental drivers are such that the past may no longer be a reasonable indicator of the future. Further research will help reduce these uncertainties.

Key hazards to the livestock sector relate not only to climate change trends but also, and importantly, to climate variability and climate extreme events such as heat waves, droughts, floods, cyclones, and wildfires. Increases in inter-annual variability of forage availability is especially a concern for the grazing sector. The increase in frequency, intensity and duration of heat waves also poses a major threat to animal health and human labour where access to mechanization and cooling systems is limited. These climate-related hazards can exacerbate other stressors with negative impacts, especially for people living in poverty. The nature and extent of such impacts are however still largely unknown and warrant further research.

In the face of global warming, and overall harmful impacts as highlighted in this review, the existing suite of adaptation strategies across ecological, socioeconomic, and institutional systems and coping range that have been developed in response to existing weather patterns may not be enough. More transformative climate adaptation may be required. These range from farm management adjustments, technological developments, income-related responses to institutional changes and, in extreme cases, abandonment of livestock keeping (see reviews by Escarcha et al. (2018), Henry et al. (2018), Herrero et al. (2016, 2020), Rivera-Ferre et al. (2016), Salman et al. (2019), Sejian et al. (2015) and Weindl et al. (2015)). Livestock ownership may also be valued as a climate risk-mitigation strategy in some contexts. Indeed, livestock plays multiple roles beyond producing food for the market, especially in smallholder farming systems. For instance, animals are capital assets that can contribute to income diversification. Livestock are also often more resilient than crops to high climate variability.

Adaptation choices and risk management actions across temporal and spatial scales and contexts will need to build on robust methods of designing, implementing and evaluating detailed development pathways. Such pathways, yet to be fully elucidated, must strengthen climate-resilience and limit trade-offs between different actors. Choices and actions will need to account for the widest possible range of potential impacts, including those with low probability but large consequences, because of the large future uncertainties related to hazards, exposure and vulnerabilities, and the large potential consequences for the sector.

Declaration of competing interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

The inputs of C.M.G, M.H and D.M contribute to the Bill and Melinda Gates Foundation Supporting Evidence Based Interventions - LiveGAPS 2 project (OPP1134229). The inputs of C.M.G. were also supported by the Army Research Office under award no. W911NF1910013 as part of the Defense Advanced Research Project Agency (DARPA) World Modelers program. The views and conclusions contained in this manuscript are those of the authors and should not be interpreted as representing the official policies either expressed or implied of the Army Research Office or the US Government. P.K.T acknowledges support provided to the CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS) from CGIAR Fund Donors, and through bilateral funding agreements (please see ccafs.cgiar.org/donors).

Footnotes

Appendix ASupplementary data to this article can be found online at https://doi.org/10.1016/j.gfs.2020.100488.

Appendix A. Supplementary data

The following is the Supplementary data to this article:

Multimedia component 1:

Click here to view.(168K, docx)Multimedia component 1

References

Abrol D.P. Pollination Biology: Biodiversity Conservation and Agricultural Production. Springer; 2011. Climate change and pollinators; pp. 479–508. [CrossRef] [Google Scholar]

Ahmed B.M.S., Younas U., Asar T.O., Dikmen S., Hansen P.J., Dahl G.E. Cows exposed to heat stress during fetal life exhibit improved thermal tolerance. J. Anim. Sci. 2017;95:3497–3503. doi:10.2527/jas2016.1298. [PubMed] [CrossRef] [Google Scholar]

Ainsworth E.A., Lemonnier P., Wedow J.M. The influence of rising tropospheric carbon dioxide and ozone on plant productivity. Plant Biol. 2020;22:5–11. doi:10.1111/plb.12973. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Alam M.Z., Carpenter-Boggs L., Mitra S., Haque M.M., Halsey J., Rokonuzzaman M., Saha B., Moniruzzaman M. Effect of salinity intrusion on food crops, livestock, and fish species at kalapara coastal belt in Bangladesh. Hindawi J. Food Qual. 2017:1–23. doi:10.1155/2017/2045157. [CrossRef] [Google Scholar]

Alary V., Aboul-Naga A., Shafie M. El, Abdelkrim N., Hamdon H., Metawi H. Roles of small ruminants in rural livelihood improvement – comparative analysis in Egypt. Rev. Elev. Med. Vet. Pays Trop. 2015;68:79–85. [Google Scholar]

Aldrich C.G., Paterson J.A., Tate J.L., Kerley M.S. The effects of endophyte-infected tall fescue consumption on diet utilization and thermal regulation in cattle. J. Anim. Sci. 1993;71:164–170. doi:10.2527/1993.711164x. [PubMed] [CrossRef] [Google Scholar]

Amstislavski P., Zubov L., Chen H., Ceccato P., Pekel J.F., Weedon J. Effects of increase in temperature and open water on transmigration and access to health care by the Nenets reindeer herders in northern Russia. Int. J. Circumpolar Health. 2013;72:21183. doi:10.3402/ijch.v72i0.21183. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Archer S.R., Andersen E.M., Predick K.I., Schwinning S., Steidl R.J., Woods S.R. Woody plant encroachment: causes and consequences. In: Briske D.D., editor. Rangeland Systems. Springer Series on Environmental Management; Cham, Switzerland: 2017. pp. 25–84. [CrossRef] [Google Scholar]

Archibald J. Disease in the dust: experiences of Q fever during drought in Australia. Perspect. Public Health. 2019;139:77–78. doi:10.1177/1757913918823423. [PubMed] [CrossRef] [Google Scholar]

Asseng S., Ewert F., Martre P., Rötter R.P., Lobell D.B., Cammarano D., Kimball B.A., Ottman M.J., Wall G.W., White J.W., Reynolds M.P., Alderman P.D., Prasad P.V.V., Aggarwal P.K., Anothai J., Basso B., Biernath C., Challinor A.J., De Sanctis G., Doltra J., Fereres E., Garcia-Vila M., Gayler S., Hoogenboom G., Hunt L.A., Izaurralde R.C., Jabloun M., Jones C.D., Kersebaum K.C., Koehler A.K., Müller C., Naresh Kumar S., Nendel C., O’leary G., Olesen J.E., Palosuo T., Priesack E., Eyshi Rezaei E., Ruane A.C., sem*nov M.A., Shcherbak I., Stöckle C., Stratonovitch P., Streck T., Supit I., Tao F., Thorburn P.J., Waha K., Wang E., Wallach D., Wolf J., Zhao Z., Zhu Y. Rising temperatures reduce global wheat production. Nat. Clim. Change. 2015;5:143–147. doi:10.1038/nclimate2470. [CrossRef] [Google Scholar]

Attavanich W., McCarl B.A., Ahmedov Z., Fuller S.W., Vedenov D.V. Effects of climate change on US grain transport. Nat. Clim. Change. 2013;3:638–643. doi:10.1038/nclimate1892. [CrossRef] [Google Scholar]

Augustine D.J., Blumenthal D.M., Springer T.L., LeCain D.R., Gunter S.A., Derner J.D. Elevated CO2 induces substantial and persistent declines in forage quality irrespective of warming in mixedgrass prairie. Ecol. Appl. 2018;28:721–735. doi:10.1002/eap.1680. [PubMed] [CrossRef] [Google Scholar]

Australian Department of Agriculture and Water Resources . 2019. Australian Meat and Live-Stock Industry ( Prohibition of Export of Sheep by Sea to Middle East — Northern Summer) Order 2019.https://www.legislation.gov.au/Details/F2019L00501 [Google Scholar]

Bagath M., Krishnan G., Devaraj C., Rashamol V.P., Pragna P., Lees A.M., Sejian V. The impact of heat stress on the immune system in dairy cattle: a review. Res. Vet. Sci. 2019;126:94–102. doi:10.1016/j.rvsc.2019.08.011. [PubMed] [CrossRef] [Google Scholar]

Bahadur Kc K., Pant L.P., Fraser E.D.G., Shrestha P.K., Shrestha D., Lama A. Assessing links between crop diversity and food self-sufficiency in three agroecological regions of Nepal. Reg. Environ. Change. 2016;16:1239–1251. doi:10.1007/s10113-015-0851-9. [CrossRef] [Google Scholar]

Bale J.S., Masters G.J., Hodkinson I.D., Awmack C., Bezemer T.M., Brown V.K., Butterfield J., Buse A., Coulson J.C., Farrar J., Good J.E.G., Harrington R., Hartley S., Jones T.H., Lindroth R.L., Press M.C., Symrnioudis I., Watt A.D., Whittaker J.B. Herbivory in global climate change research: direct effects of rising temperature on insect herbivores. Global Change Biol. 2002;8:1–16. doi:10.1046/j.1365-2486.2002.00451.x. [CrossRef] [Google Scholar]

Barbehenn R.V., Chen Z., Karowe D.N., Spickard A. C3 grasses have higher nutritional quality than C4 grasses under ambient and elevated atmospheric CO2. Global Change Biol. 2004;10:1565–1575. doi:10.1111/j.1365-2486.2004.00833.x. [CrossRef] [Google Scholar]

BBC Coronavirus: meat shortage leaves US farmers with “mind-blowing. 2020. choice [WWW Document]. URL https://www.bbc.com/news/world-us-canada-52575904.

Bebber D.P., Ramotowski M.A.T., Gurr S.J. Crop pests and pathogens move polewards in a warming world. Nat. Clim. Change. 2013;3:985–988. doi:10.1038/nclimate1990. [CrossRef] [Google Scholar]

Bernabucci U., Lacetera N., Baumgard L.H., Rhoads R.P., Ronchi B., Nardone A. Metabolic and hormonal acclimation to heat stress in domesticated ruminants. Animal. 2010;4:1167–1183. doi:10.1017/S175173111000090X. [PubMed] [CrossRef] [Google Scholar]

Bernabucci U., Lacetera N., Ronchi B., Nardone A. Effects of the hot season on milk protein fractions in Holstein cows. Anim. Res. 2002;51:25–33. doi:10.1051/animres:2002006. [CrossRef] [Google Scholar]

Biswas B., Qi F., Biswas J., Wijayawardena A., Khan M., Naidu R. The fate of chemical pollutants with soil properties and processes in the climate change paradigm—a review. Soil Syst. 2018;2:51. doi:10.3390/soilsystems2030051. [CrossRef] [Google Scholar]

Boddicker R.L., Seibert J.T., Johnson J.S., Pearce S.C., Selsby J.T., Gabler N.K., Lucy M.C., Safranski T.J., Rhoads R.P., Baumgard L.H., Ross J.W. Gestational heat stress alters postnatal offspring body composition indices and metabolic parameters in pigs. PloS One. 2014;9 doi:10.1371/journal.pone.0110859. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Bortolussi G., McIvor J.G., Hodgkinson J.J., Coffey S.G., Holmes C.R. The northern Australian beef industry, a snapshot. 1. Regional enterprise activity and structure. Aust. J. Exp. Agric. 2005;45:1057–1073. doi:10.1071/EA03096. [CrossRef] [Google Scholar]

Bourguignon M., Nelson J.A., Carlisle E., Ji H., Dinkins R.D., Phillips T.D., McCulley R.L. Ecophysiological responses of tall fescue genotypes to fungal endophyte infection, elevated temperature, and precipitation. Crop Sci. 2015;55:2895–2909. doi:10.2135/cropsci2015.01.0020. [CrossRef] [Google Scholar]

Braide W., Justice-Alucho C.H., Ohabughiro N., Adeleye S.A. Global climate change and changes in disease distribution: a review in retrospect. Int. J. Adv. Res. Biol. Sci. 2020;7:32–46. doi:10.22192/ijarbs. [CrossRef] [Google Scholar]

Brown A.L., Cavagnaro T.R., Gleadow R., Miller R.E. Interactive effects of temperature and drought on cassava growth and toxicity: implications for food security? Global Change Biol. 2016;22:3461–3473. doi:10.1111/gcb.13380. [PubMed] [CrossRef] [Google Scholar]

Burke M., Hsiang S.M., Miguel E. Global non-linear effect of temperature on economic production. Nature. 2015;527:235–239. doi:10.1038/nature15725. [PubMed] [CrossRef] [Google Scholar]

Caminade C., McIntyre K.M., Jones A.E. Impact of recent and future climate change on vector-borne diseases. Ann. N. Y. Acad. Sci. 2019;1436:157–173. doi:10.1111/nyas.13950. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Carlson R.E. Heat stress, plant-available soil moisture, and corn yields in Iowa: a short- and long-term view. J. Prod. Agric. 1990;3:293–297. doi:10.2134/jpa1990.0293. [CrossRef] [Google Scholar]

Caulfield M.P., Cambridge H., Foster S.F., McGreevy P.D. Heat stress: a major contributor to poor animal welfare associated with long-haul live export voyages. Vet. J. 2014;199:223–228. doi:10.1016/j.tvjl.2013.09.018. [PubMed] [CrossRef] [Google Scholar]

Chang J., Ciais P., Viovy N., Soussana J.F., Klumpp K., Sultan B. Future productivity and phenology changes in European grasslands for different warming levels: implications for grassland management and carbon balance. Carbon Bal. Manag. 2017;12 doi:10.1186/s13021-017-0079-8. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Changnon S.A. The 1988 drought, barges, and diversion. Bull. Am. Meteorol. Soc. 1989;70:1092–1104. doi:10.1175/1520-0477(1989)070<1092:TDBAD>2.0.CO;2. [CrossRef] [Google Scholar]

Christodouloum A., Demirel H. Impacts of climate change on transport. 2017. Luxembourg. Joint Research Centre (European Commission) [CrossRef]

CIRAD . CIRAD; Montpellier: 2016. Livestock Farming & Local Development. [Google Scholar]

Clements K.W., Si J.W. Engel's law, diet diversity, and the quality of food consumption. Am. J. Agric. Econ. 2018;100:1–22. doi:10.1093/ajae/aax053. [CrossRef] [Google Scholar]

Collins T., Hampton J.O., Barnes A.L. A systematic review of heat load in australian livestock transported by sea. Animals. 2018;8:164. doi:10.3390/ani8100164. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Coulomb D. Refrigeration and cold chain serving the global food industry and creating a better future: two key IIR challenges for improved health and environment. Trends Food Sci. Technol. 2008;19:413–417. doi:10.1016/j.tifs.2008.03.006. [CrossRef] [Google Scholar]

Craine J.M., Nippert J.B., Elmore A.J., Skibbe A.M., Hutchinson S.L., Brunsell N.A. Timing of climate variability and grassland productivity. Proc. Natl. Acad. Sci. Unit. States Am. 2012;109:3401–3405. doi:10.1073/pnas.1118438109. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Crouse J.D., Cundiff L.V., Koch R.M., Koohmaraie M., Seideman S.C. Comparisons of and inheritance for carcass beef characteristics and meat palatability. J. Anim. Sci. 1989;67:2661–2668. doi:10.2527/jas1989.67102661x. [CrossRef] [Google Scholar]

D'Souza R.M., Becker N.G., Hall G., Moodie K.B.A. Does ambient temperature affect foodborne disease? Epidemiology. 2004;15:86–92. doi:10.1097/01.ede.0000101021.03453.3e. [PubMed] [CrossRef] [Google Scholar]

Dahl G.E., Tao S., Monteiro A.P.A. Effects of late-gestation heat stress on immunity and performance of calves. J. Dairy Sci. 2016;99:3193–3198. doi:10.3168/jds.2015-9990. [PubMed] [CrossRef] [Google Scholar]

Dalintai N.G., Gauwau N., Yanbo L., Enkhee J., Shurun L. The new Otor: risk management in a desert grassland. In: Fernández-Giménez M.E., Wang X., Batkhishig B., Klein J.A., Reid R.S., editors. Restoring Community Connections to the Land. CAB International; Wallingford, UK: 2012. [Google Scholar]

Dearing M.D. Temperature-dependent toxicity in mammals with implications for herbivores: a review. J. Comp. Physiol. B Biochem. Syst. Environ. Physiol. 2013;183:43–50. doi:10.1007/s00360-012-0670-y. [PubMed] [CrossRef] [Google Scholar]

Dube O.P., Pickup G. Effects of rainfall variability and communal and semi- commercial grazing on land cover in southern African rangelands. Clim. Res. 2001;17:195–208. doi:10.3354/cr017195. [CrossRef] [Google Scholar]

Egberts V., van Schaik G., Brunekreef B., Hoek G. Short-term effects of air pollution and temperature on cattle mortality in The Netherlands. Prev. Vet. Med. 2019;168:1–8. doi:10.1016/j.prevetmed.2019.03.021. [PubMed] [CrossRef] [Google Scholar]

Ellis F. Oxford University Press; Oxford: 2000. Rural Livelihoods and Diversity in Developing Countries. [Google Scholar]

Escarcha J.F., Lassa J.A., Zander K.K. Livestock under climate change: a systematic review of impacts and adaptation. Climate. 2018;6:1–17. doi:10.3390/cli6030054. [CrossRef] [Google Scholar]

European Commission . 2009. The Evolution of Value-Added Repartition along the European Food Supply Chain.https://ec.europa.eu/economy_finance/publications/pages/publication16075_en.pdf (Brussels) [Google Scholar]

Eyring V., Cox P.M., Flato G.M., Gleckler P.J., Abramowitz G., Caldwell P., Collins W.D., Gier B.K., Hall A.D., Hoffman F.M., Hurtt G.C., Jahn A., Jones C.D., Klein S.A., Krasting J.P., Kwiatkowski L., Lorenz R., Maloney E., Meehl G.A., Pendergrass A.G., Pincus R., Ruane A.C., Russell J.L., Sanderson B.M., Santer B.D., Sherwood S.C., Simpson I.R., Stouffer R.J., Williamson M.S. Taking climate model evaluation to the next level. Nat. Clim. Change. 2019;9:102–110. doi:10.1038/s41558-018-0355-y. [CrossRef] [Google Scholar]

FAO I.F.A.D., UNICEF W. and W. FAO; Rome: 2018. The State of Food Security and Nutrition in the World 2018. Building Climate Resilience for Food Security and Nutrition. [CrossRef] [Google Scholar]

FAO . 2020. Impact of Desert Locust Infestation on Household Livelihoods and Food Security in Ethiopia - Joint Assessment Findings.https://reliefweb.int/report/ethiopia/impact-desert-locust-infestation-household-livelihoods-and-food-security-ethiopia [Google Scholar]

FAO Food security indicators. 2019. http://www.fao.org/economic/ess/ess-fs/ess-fadata#.XtnkLUFS91M [WWW Document]. URL.

FAO . Guidelines for assessment. Livestock Environmental Assessment and Performance (LEAP) Partnership; Rome: 2019. Water Use in Livestock Production Systems and Supply Chains – Guidelines for Assessment (Version 1), Water Use in Livestock Production Systems and Supply Chains. [CrossRef] [Google Scholar]

FAO Global livestock environmental assessment model. 2018. http://www.fao.org/gleam/results/en/ (GLEAM) [WWW Document]. URL. [PubMed]

FAO . 2017. Global Livestock Environmental Assessment Model. Model Description.http://www.fao.org/fileadmin/user_upload/gleam/docs/GLEAM_2.0_Model_description.pdf Version 2.0. Rome, Italy. [Google Scholar]

FAO . 2014. Youth and Agriculture.http://www.fao.org/3/a-i3947e.pdf Rome. [Google Scholar]

FAOSTAT Faostat. 2020. http://www.fao.org/faostat/en/ WWW Document]. URL.

Faye B., Chaibou M., Vias G. Integrated impact of climate change and socioeconomic development on the evolution of camel farming systems. Br. J. Environ. Clim. Change. 2012;2:227–244. doi:10.9734/bjecc/2012/1548. [CrossRef] [Google Scholar]

Filipe J.F., Herrera V., Curone G., Vigo D., Riva F. Floods, hurricanes, and other catastrophes: a challenge for the immune system of livestock and other animals. Front. Vet. Sci. 2020;7:1–8. doi:10.3389/fvets.2020.00016. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Flood J. The importance of plant health to food security. Food Secur. 2010;2:215–231. doi:10.1007/s12571-010-0072-5. [CrossRef] [Google Scholar]

Frimpong K., Eddie Van Etten E.J., Oosthuzien J., Fannam Nunfam V. Heat exposure on farmers in northeast Ghana. Int. J. Biometeorol. 2017;61:397–406. doi:10.1007/s00484-016-1219-7. [PubMed] [CrossRef] [Google Scholar]

Gaughan J.B., Mader T.L., Holt S.M., Lisle A. A new heat load index for feedlot cattle. J. Anim. Sci. 2008;86:226–234. doi:10.2527/jas.2007-0305. [PubMed] [CrossRef] [Google Scholar]

Gauly M., Ammer S. Review: challenges for dairy cow production systems arising from climate changes. Animal. 2020;14:S196–S203. doi:10.1017/S1751731119003239. [PubMed] [CrossRef] [Google Scholar]

Georgia Department of Agriculture Press release - Hurricane Michael Devastates Georgia's agricultural industry. 2018. http://agr.georgia.gov/hurricane-michael-devastates-georgias-agricultural-industry.aspx WWW Document]. URL.

Gleadow R.M., Evans J.R., Mccaffery S., Cavagnaro T.R. Growth and nutritive value of cassava (Manihot esculenta Cranz.) are reduced when grown in elevated CO2. Plant Biol. 2009;11:76–82. doi:10.1111/j.1438-8677.2009.00238.x. [PubMed] [CrossRef] [Google Scholar]

Gleadow R.M., Ottman M.J., Kimball B.A., Wall G.W., Pinter P.J., LaMorte R.L., Leavitt S.W. Drought-induced changes in nitrogen partitioning between cyanide and nitrate in leaves and stems of sorghum grown at elevated CO2 are age dependent. Field Crop. Res. 2016;185:97–102. doi:10.1016/j.fcr.2015.10.010. [CrossRef] [Google Scholar]

Godber O.F., Wall R. Livestock and food security: vulnerability to population growth and climate change. Global Change Biol. 2014;20:3092–3102. doi:10.1111/gcb.12589. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Godde C.M., Boone R., Ash A.J., Waha K., Sloat L., Thornton P.K., Herrero M. Global rangeland production systems and livelihoods at threat under climate change and variability. Environ. Res. Lett. 2020;15:44021. doi:10.1088/1748-9326/ab7395. [CrossRef] [Google Scholar]

Godfray H.C.J., Aveyard P., Garnett T., Hall J.W., Key T.J., Lorimer J., Pierrehumbert R.T., Scarborough P., Springmann M., Jebb S.A. Meat consumption, health, and the environment. Science. 2018;80:243. doi:10.1126/science.aam5324. [PubMed] [CrossRef] [Google Scholar]

Gonzalez-Rivas P.A., Chauhan S.S., Ha M., Fegan N., Dunshea F.R., Warner R.D. Effects of heat stress on animal physiology, metabolism, and meat quality: a review. Meat Sci. 2020;162:108025. doi:10.1016/j.meatsci.2019.108025. [PubMed] [CrossRef] [Google Scholar]

Guan K., Good S.P., Caylor K.K., Sato H., Wood E.F., Li H. Continental-scale impacts of intra-seasonal rainfall variability on simulated ecosystem responses in Africa. Biogeosciences. 2014;11:6939–6954. doi:10.5194/bg-11-6939-2014. [CrossRef] [Google Scholar]

Hallegatte S., Rozenberg J. Climate change through a poverty lens. Nat. Clim. Change. 2017;7:250–256. doi:10.1038/nclimate3253. [CrossRef] [Google Scholar]

Hasegawa T., Fujimori S., Havlík P., Valin H., Bodirsky B.L., Doelman J.C., Fellmann T., Kyle P., Koopman J.F.L., Lotze-Campen H., Mason-D’Croz D., Ochi Y., Pérez Domínguez I., Stehfest E., Sulser T.B., Tabeau A., Takahashi K., Takakura J., van Meijl H., van Zeist W.-J., Wiebe K., Witzke P. Risk of increased food insecurity under stringent global climate change mitigation policy. Nat. Clim. Change. 2018;8:699–703. doi:10.1038/s41558-018-0230-x. [CrossRef] [Google Scholar]

Hatfield J.L., Boote K.J., Kimball B.A., Ziska L.H., Izaurralde R.C., Ort D., Thomson A.M., Wolfe D. Climate impacts on agriculture: implications for crop production. Agron. J. 2011;103:351–370. doi:10.2134/agronj2010.0303. [CrossRef] [Google Scholar]

Havlík P., Leclère D., Valin H., Herrero M., Schmid E., Soussana J., Obersteiner M. Global climate change, food supply and livestock production systems: a bioeconomic analysis. In: Elbehri A., editor. Climate Change and Food Systems: Global Assessments and Implications for Food Security and Trade. Food Agriculture Organization of the United Nations (FAO); Rome: 2015. pp. 178–197. ISBN 978-92-5-108699-5. [Google Scholar]

Headey D., Fan S. Research Report of the International Food Policy Research Institute; 2010. Reflections on the global food crisis: how did it happen? how has it hurt? and how can we prevent the next one? [CrossRef] [Google Scholar]

Hegland S.J., Nielsen A., Lazaro A., Bjerknes A.L., Totland O. How does climate warming affect plant-pollinator interactions? Ecol. Lett. 2009;12:184–195. doi:10.1111/j.1461-0248.2008.01269.x. [PubMed] [CrossRef] [Google Scholar]

Henry B.K., Eckard R.J., Beauchemin K.A. Review: adaptation of ruminant livestock production systems to climate changes. Animal. 2018;12:S445–S456. doi:10.1017/S1751731118001301. [PubMed] [CrossRef] [Google Scholar]

Herrero M., Addison J., Bedelian C., Carabine E., Havlík P., Henderson B., Steeg J. Van De, Thornton P.K. Climate change and pastoralism: impacts, consequences and adaptation. Rev. Sci. Tech. Off. Int. Epiz. 2016;35 https://web.oie.int/boutique/extrait/07herrero417433.pdf [PubMed] [Google Scholar]

Herrero M., Havlík P., Valin H., Notenbaert A.M., Rufino M.C., Thornton P.K., Blümmel M., Weiss F., Grace D., Obersteiner M., Havlíkb P., Valin H., Notenbaert A.M., Rufino M.C., Thornton P.K., Blümmel M., Weiss F., Grace D., Obersteiner M. Biomass use, production, feed efficiencies, and greenhouse gas emissions from global livestock systems. Proc. Natl. Acad. Sci. U. S. A. 2013;110:20888. doi:10.1073/pnas.1308149110. –93. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Herrero M., Mason-D’Croz D., Godde C.M., Palmer J., Thornton P.K., Gill M. CGIAR Science Forum 2018. CGIAR; Stellenbosch, South Africa: 2018. Livestock, land and the environmental limits of animal source-food consumption; pp. 1–39. [Google Scholar]

Hirakawa R., Nurjanah S., Furukawa K., Murai A., Kikusato M., Nochi T., Toyomizu M. Heat stress causes immune abnormalities via massive damage to effect proliferation and differentiation of lymphocytes in broiler chickens. Front. Vet. Sci. 2020;7:1–13. doi:10.3389/fvets.2020.00046. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Hobbs N.T., Galvin K.A., Stokes C.J., Lackett J.M., Ash A.J., Boone R.B., Reid R.S., Thornton P.K. Fragmentation of rangelands: implications for humans, animals, and landscapes. Global Environ. Change. 2008;18:776–785. doi:10.1016/j.gloenvcha.2008.07.011. [CrossRef] [Google Scholar]

Hodges R.J., Buzby J.C., Bennett B. Postharvest losses and waste in developed and less developed countries: opportunities to improve resource use. J. Agric. Sci. 2011;149:37–45. doi:10.1017/S0021859610000936. [CrossRef] [Google Scholar]

Hristov A.N., Degaetano A.T., Rotz C.A., Hoberg E., Skinner R.H., Felix T., Li H., Patterson P.H., Roth G., Hall M., Ott T.L., Baumgard L.H., Staniar W., Hulet R.M., Dell C.J., Brito A.F., Hollinger D.Y. Climate change effects on livestock in the Northeast US and strategies for adaptation. Climatic Change. 2018;146:33–45. doi:10.1007/s10584-017-2023-z. [CrossRef] [Google Scholar]

Hruska T., Huntsinger L., Brunson M., Li W., Marshall N., Oviedo J.L., Whitcomb H. Rangelands as Social–Ecological Systems. In: Briske D.D., editor. Rangeland Systems. Springer; Cham, Switzerland: 2017. pp. 263–302. [CrossRef] [Google Scholar]

Huang J., Li Y., Fu C., Chen F., Fu Q., Dai A., Shinoda M., Ma Z., Guo W., Li Z., Zhang L., Liu Y., Yu H., He Y., Xie Y., Guan X., Ji M., Lin L., Wang S., Yan H., Wang G. Dryland climate change: recent progress and challenges. Rev. Geophys. 2017;55:719–778. doi:10.1002/2016RG000550. [CrossRef] [Google Scholar]

Humphrey J. 2017. IFAD RESEARCH SERIES 11 - Food Safety, Trade, Standards and the Integration of Smallholders into Value Chains: a Review of the Literature. [Google Scholar]

International Institute of Refrigeration . Informatory Note on Refrigeration and Food; Paris: 2009. The Role of Refrigeration in Worldwide Nutrition (No. 5th) [Google Scholar]

International Labour Organization . 2019. Working on a Warmer Planet: the Impact of Heat Stress on Labour Productivity and Decent Work.https://www.ilo.org/global/publications/books/WCMS_711919/lang--en/index.htm Geneva. [Google Scholar]

IPBES . Secretariat of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services; Bonn, Germany: 2016. IPBES (2016): Summary for Policymakers of the Assessment Report of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services on Pollinators, Pollination and Food Production.https://ipbes.net/assessment-reports/pollinators [Google Scholar]

IPCC . 2014. Climate Change 2014 Part A: Global and Sectoral Aspects, Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change.https://www.ipcc.ch/report/ar5/wg2/ [Google Scholar]

IPCC . Climate change 2014: impacts, adaptation, and vulnerability - summary for policy makers. In: Field C.B., Barros V.R., Dokken D.J., Mach K.J., Mastrandrea M.D., Bilir T.E., Chatterjee M., Ebi K.L., Estrada Y.O., Genova R.C., Girma B., Kissel E.S., Levy A.N., MacCracken S., Mastrandrea P.R., White L.L., editors. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press; Cambridge, United Kingdom and New York, NY, USA: 2014. pp. 1–32.https://www.ipcc.ch/report/ar5/wg2/summary-for-policymakers/ [Google Scholar]

IPCC . Ipcc; 2014. Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Google Scholar]

Islam S.N., Winkel J. Climate change and social inequality (No. 152) 2017. https://www.un.org/esa/desa/papers/2017/wp152_2017.pdf New York.

Izaurralde R.C., Thomson A.M., Morgan J.A., Fay P.A., Polley H.W., Hatfield J.L. Climate impacts on agriculture: implications for forage and rangeland production. Agron. J. 2011;103:371–381. doi:10.2134/agronj2010.0304. [CrossRef] [Google Scholar]

James S.J., James C. The food cold-chain and climate change. Food Res. Int. 2010;43:1944–1956. doi:10.1016/j.foodres.2010.02.001. [CrossRef] [Google Scholar]

Jiménez Cisneros B.E., Oki T., Arnell N.W., Benito G., Cogley J.G., Döll P., Jiang T., Mwakalila S.S. Freshwater resources. In: Field C.B., Barros V.R., Dokken D.J., Mach K.J., Mastrandrea M.D., Bilir T.E., Chatterjee M., Ebi K.L., Estrada Y.O., Genova R.C., Girma B., Kissel E.S., Levy A.N., MacCracken S., Mastrandrea P.R., White L.L., editors. Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press; Cambridge, United Kingdom and New York, NY, USA: 2014. pp. 229–269. [Google Scholar]

Johnson D.D., Huffman R.D., Williams S.E., Hargrove D.D. Effects of percentage Brahman and Angus breeding, age-season of feeding and slaughter end point on meat palatability and muscle characteristics. J. Anim. Sci. 1990;68:1980–1986. doi:10.2527/1990.6871980x. [PubMed] [CrossRef] [Google Scholar]

Johnson J.S., Sanz Fernandez M.V., Patience J.F., Ross J.W., Gabler N.K., Lucy M.C., Safranski T.J., Rhoads R.P., Baumgard L.H. Effects of in utero heat stress on postnatal body composition in pigs: II. Finishing phase. J. Anim. Sci. 2015;93:82–92. https://academic.oup.com/jas/article/93/1/82/4701525 [PubMed] [Google Scholar]

Jones A.D. The production diversity of subsistence farms in the Bolivian Andes is associated with the quality of child feeding practices as measured by a validated summary feeding index. Publ. Health Nutr. 2015;18:329–342. doi:10.1017/S1368980014000123. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Jones P.G., Thornton P.K. Croppers to livestock keepers: livelihood transitions to 2050 in Africa due to climate change. Environ. Sci. Pol. 2009;12:427–437. doi:10.1016/j.envsci.2008.08.006. [CrossRef] [Google Scholar]

Kagunyu A.W., Wanjohi J. Camel rearing replacing cattle production among the Borana community in Isiolo County of Northern Kenya, as climate variability bites. Pastoralism. 2014;4:1–5. doi:10.1186/s13570-014-0013-6. [CrossRef] [Google Scholar]

King M., Altdorff D., Li P., Galagedara L., Holden J., Unc A. Northward shift of the agricultural climate zone under 21st-century global climate change. Sci. Rep. 2018;8:7904. doi:10.1038/s41598-018-26321-8. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Kjellstrom T., Briggs D., Freyberg C., Lemke B., Otto M., Hyatt O. Heat, human performance, and occupational health: a key issue for the assessment of global climate change impacts. Annu. Rev. Publ. Health. 2016;37:97–112. doi:10.1146/annurev-publhealth-032315-021740. [PubMed] [CrossRef] [Google Scholar]

Kjellstrom T., Kovats R.S., Lloyd S.J., Holt T., Tol R.S.J. The direct impact of climate change on regional labor productivity. Arch. Environ. Occup. Health. 2009;64:217–227. doi:10.1080/19338240903352776. [PubMed] [CrossRef] [Google Scholar]

Klein A.M., Vaissière B.E., Cane J.H., Steffan-Dewenter I., Cunningham S.A., Kremen C., Tscharntke T. Importance of pollinators in changing landscapes for world crops. Proc. R. Soc. B Biol. Sci. 2007;274:303–313. doi:10.1098/rspb.2006.3721. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Knittel N., Jury M.W., Bednar-Friedl B., Bachner G., Steiner A.K. A global analysis of heat-related labour productivity losses under climate change—implications for Germany's foreign trade. Climatic Change. 2020;160:251–269. doi:10.1007/s10584-020-02661-1. [CrossRef] [Google Scholar]

Kurnath P., Merz N.D., Dearing M.D. Ambient temperature influences tolerance to plant secondary compounds in a mammalian herbivore. Proc. R. Soc. B Biol. Sci. 2016;283 doi:10.1098/rspb.2015.2387. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Lambar E.F., Thomas G. The health and well-being of North Carolina's farmworkers. N. C. Med. J. 2019;80:107–112. doi:10.18043/ncm.80.2.107. [PubMed] [CrossRef] [Google Scholar]

Lara L.J., Rostagno M.H. Impact of heat stress on poultry production. Animals. 2013;3:356–369. doi:10.3390/ani3020356. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Lau C.L., Smythe L.D., Craig S.B., Weinstein P. Climate change, flooding, urbanisation and leptospirosis: fuelling the fire? Trans. R. Soc. Trop. Med. Hyg. 2010;104:631–638. doi:10.1016/j.trstmh.2010.07.002. [PubMed] [CrossRef] [Google Scholar]

Lees A.M., Sejian V., Wallage A.L., Steel C.C., Mader T.L., Lees J.C., Gaughan J.B. The impact of heat load on cattle. Animals. 2019;9:322. doi:10.3390/ani9060322. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Legesse G., Ominski K.H., Beauchemin K.A., Pfister S., Martel M., McGeough E.J., Hoekstra A.Y., Kroebel R., Cordeiro M.R.C., McAllister T.A. BOARD-invited review: quantifying water use in ruminant production. J. Anim. Sci. 2017;95:2001–2018. https://academic.oup.com/jas/article/95/5/2001/4703559 [PubMed] [Google Scholar]

Lemonte J.J., Stuckey J.W., Sanchez J.Z., Tappero R., Rinklebe J., Sparks D.L. sea level rise induced arsenic release from historically contaminated coastal soils. Environ. Sci. Technol. 2017;51:5913–5922. doi:10.1021/acs.est.6b06152. [PubMed] [CrossRef] [Google Scholar]

Lindstad H., Bright R.M., Strømman A.H. Economic savings linked to future Arctic shipping trade are at odds with climate change mitigation. Transport Pol. 2016;45:24–30. doi:10.1016/j.tranpol.2015.09.002. [CrossRef] [Google Scholar]

Lloyds . 2015. Food System Shock: the Insurance Impacts of Acute Disruption to Global Food Supply. [Google Scholar]

Lobell D., Schlenker W., Costa-Roberts J. Climate trends and global crop production since 1980. Science (80-. ) 2011;333:616–621. https://science.sciencemag.org/content/333/6042/616 [PubMed] [Google Scholar]

Lobell D.B., Bänziger M., Magorokosho C., Vivek B. Nonlinear heat effects on African maize as evidenced by historical yield trials. Nat. Clim. Change. 2011;1:42–45. doi:10.1038/nclimate1043. [CrossRef] [Google Scholar]

Lobell D.B., Hammer G.L., McLean G., Messina C., Roberts M.J., Schlenker W. The critical role of extreme heat for maize production in the United States. Nat. Clim. Change. 2013;3:497–501. doi:10.1038/nclimate1832. [CrossRef] [Google Scholar]

Loladze I. Hidden shift of the ionome of plants exposed to elevated CO2 depletes minerals at the base of human nutrition. Elife. 2014;1–29 doi:10.7554/eLife.02245. 2014. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Ludlow M. Stress physiology of tropical pasture plants. Trop. grasslands. 1980;14:136–145. [Google Scholar]

Markolf S.A., Hoehne C., Fraser A., Chester M.V., Underwood B.S. Transportation resilience to climate change and extreme weather events – beyond risk and robustness. Transport Pol. 2019;74:174–186. doi:10.1016/j.tranpol.2018.11.003. [CrossRef] [Google Scholar]

Marrugo-Negrete J., Pinedo-Hernández J., Combatt E.M., Bravo A.G., Díez S. Flood-induced metal contamination in the topsoil of floodplain agricultural soils: a case-study in Colombia. Land Degrad. Dev. 2019;30:2139–2149. doi:10.1002/ldr.3398. [CrossRef] [Google Scholar]

Marshall N.A. Adaptive capacity on the northern Australian rangelands. Rangel. J. 2015;37:617–622. doi:10.1071/RJ15054. [CrossRef] [Google Scholar]

Mashaly M.M., Hendricks G.L., Kalama M.A., Gehad A.E., Abbas A.O., Patterson P.H. Effect of heat stress on production parameters and immune responses of commercial laying hens. Poultry Sci. 2004;83:889–894. doi:10.1093/ps/83.6.889. [PubMed] [CrossRef] [Google Scholar]

Masters D., Edwards N., Sillence M., Avery A., Revell D., Friend M., Sanford P., Saul G., Beverly C., Young J. The role of livestock in the management of dryland salinity. Aust. J. Exp. Agric. 2006;46:733–741. doi:10.1071/EA06017. [CrossRef] [Google Scholar]

Mbow C., Rosenzweig C., Barioni L.G., Benton T.G., Herrero M., Krishnapillai M., Liwenga E., Pradhan P., Rivera-Ferre M.G., Sapkota T., Tubiello F.N., Xu Y. Food security. In: Shukla P.R., Skea J., Buendia E.C., Masson-Delmotte V., Pörtner H.-O., Roberts D.C., Zhai P., Slade R., Connors S., Diemen R. van, Ferrat M., Haughey E., Luz S., Neogi S., Pathak M., Petzold J., Pereira J.P., Vyas P., Huntley E., Kissick K., Belkacemi M., Malley J., editors. Climate Change and Land: an IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems. 2019. pp. 437–550.https://www.ipcc.ch/srccl/ [Google Scholar]

McAllister R.R.J. Livestock mobility in arid and semiarid Australia: escaping variability in space. Rangel. J. 2012;34:139–147. doi:10.1071/RJ11090. [CrossRef] [Google Scholar]

McAllister R.R.J., Gordon I.J., Janssen M.A., Abel N. Pastoralists' responses to variation of rangeland resources in time and space. Ecol. Appl. 2006;16:572–583. [PubMed] [Google Scholar]

McGrath J.M., Betzelberger A.M., Wang S., Shook E., Zhu X.G., Long S.P., Ainsworth E.A. An analysis of ozone damage to historical maize and soybean yields in the United States. Proc. Natl. Acad. Sci. U. S. A. 2015;112:14390–14395. doi:10.1073/pnas.1509777112. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

McIntyre K.M., Setzkorn C., Hepworth P.J., Morand S., Morse A.P., Baylis M. Systematic assessment of the climate sensitivity of important human and domestic animals pathogens in Europe. Sci. Rep. 2017;7:1–10. doi:10.1038/s41598-017-06948-9. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Meerburg B.G., Verhagen A., Jongschaap R.E.E., Franke A.C., Schaap B.F., Dueck T.A., Van Der Werf A. Do nonlinear temperature effects indicate severe damages to US crop yields under climate change? Proc. Natl. Acad. Sci. U. S. A. 2009;106:15594–15598. doi:10.1073/pnas.0910618106. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Mekonnen M.M., Hoekstra A.Y. A global assessment of the water footprint of farm animal products. Ecosystems. 2012;15:401–415. doi:10.1007/s10021-011-9517-8. [CrossRef] [Google Scholar]

Memmott J., Craze P.G., Waser N.M., Price M.V. Global warming and the disruption of plant-pollinator interactions. Ecol. Lett. 2007;10:710–717. doi:10.1111/j.1461-0248.2007.01061.x. [PubMed] [CrossRef] [Google Scholar]

Menzel D.B. Ozone: an overview of its toxicity in man and animals. J. Toxicol. Environ. Health. 1984;13:183–204. doi:10.1080/15287398409530493. [PubMed] [CrossRef] [Google Scholar]

Miller A.W., Ruiz G.M. Arctic shipping and marine invaders. Nat. Clim. Change. 2014;4:413–416. doi:10.1038/nclimate2244. [CrossRef] [Google Scholar]

Mills G., Pleijel H., Malley C.S., Sinha B., Cooper O.R., Schultz Martin G., Neufeld H.S., Simpson D., Sharps K., Feng Z., Gerosa G., Harmens H., Kobayashi K., Saxena P., Paoletti E., Sinha V., Xu X. Tropospheric Ozone Assessment Report: present-day tropospheric ozone distribution and trends relevant to vegetation. Elem Sci Anth. 2018;6:47. doi:10.1525/elementa.302. [CrossRef] [Google Scholar]

Monteiro A.P.A., Tao S., Thompson I.M.T., Dahl G.E. In utero heat stress decreases calf survival and performance through the first lactation. J. Dairy Sci. 2016;99:8443–8450. doi:10.3168/jds.2016-11072. [PubMed] [CrossRef] [Google Scholar]

Moore B.D., Wiggins N.L., Marsh K.J., Dearing M.D., Foley W.J. Translating physiological signals to changes in feeding behaviour in mammals and the future effects of global climate change. Anim. Prod. Sci. 2015;55:272–283. doi:10.1071/AN14487. [CrossRef] [Google Scholar]

Morrison T.H., Hettiarachchi M., Seabrook L., McAlpine C. Environmental change and social learning. In: Richardson D., Castree N., Goodchild M.F., Kobayashi A., Liu W., Marston R.A., editors. International Encyclopedia of Geography: People, the Earth, Environment and Technology. Wiley-Blackwell; Hoboken, NJ: 2017. pp. 1–4. [CrossRef] [Google Scholar]

Motoki K., Saito T., Nouchi R., Kawashima R., Sugiura M. The paradox of warmth: ambient warm temperature decreases preference for savory foods. Food Qual. Prefer. 2018;69:1–9. doi:10.1016/j.foodqual.2018.04.006. [CrossRef] [Google Scholar]

Mueller B., Seneviratne S.I. Hot days induced by precipitation deficits at the global scale. Proc. Natl. Acad. Sci. U. S. A. 2012;109:12398–12403. doi:10.1073/pnas.1204330109. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Muhammad A., D'Souza A., Meade B., Micha R., Mozaffarian D. How income and food prices influence global dietary intakes by age and sex: evidence from 164 countries. BMJ Glob. Heal. 2017;2:1–11. doi:10.1136/bmjgh-2016-000184. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Müller C., Robertson R.D. Projecting future crop productivity for global economic modeling. Agric. Econ. 2014;45:37–50. doi:10.1111/agec.12088. [CrossRef] [Google Scholar]

Murendo C., Nhau B., Mazvimavi K., Khanye T., Gwara S. Nutrition education, farm production diversity, and commercialization on household and individual dietary diversity in Zimbabwe. Food Nutr. Res. 2018;62:1276. doi:10.29219/fnr.v62.1276. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Muscat A., de Olde E.M., de Boer I.J.M., Ripoll-Bosch R. The battle for biomass: a systematic review of food-feed-fuel competition. Glob. Food Sec. 2020;25:100330. doi:10.1016/j.gfs.2019.100330. [CrossRef] [Google Scholar]

Myers S.S., Smith M.R., Guth S., Golden C.D., Vaitla B., Mueller N.D., Dangour A.D., Huybers P. Climate change and global food systems: potential impacts on food security and undernutrition. Annu. Rev. Publ. Health. 2017;38:259–277. [PubMed] [Google Scholar]

Myers S.S., Zanobetti A., Kloog I., Huybers P., Leakey A.D.B., Bloom A.J., Carlisle E., Dietterich L.H., Fitzgerald G., Hasegawa T., Holbrook N.M., Nelson R.L., Ottman M.J., Raboy V., Sakai H., Sartor K.A., Schwartz J., Seneweera S., Tausz M., Usui Y. Increasing CO2 threatens human nutrition. Nature. 2014;510:139–142. doi:10.1038/nature13179. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Nardone A., Ronchi B., Lacetera N., Ranieri M.S., Bernabucci U. Effects of climate changes on animal production and sustainability of livestock systems. Livest. Sci. 2010;130:57–69. doi:10.1016/j.livsci.2010.02.011. [CrossRef] [Google Scholar]

Nawab A., Ibtisham F., Li G., Kieser B., Wu J., Liu W., Zhao Y., Nawab Y., Li K., Xiao M., An L. Heat stress in poultry production: mitigation strategies to overcome the future challenges facing the global poultry industry. J. Therm. Biol. 2018;78:131–139. doi:10.1016/j.jtherbio.2018.08.010. [PubMed] [CrossRef] [Google Scholar]

Needs N., Environments H., Personnel M., Operations F., Marriott B.M., Isbn M., Pdf T., Academy N., Academy N., Press N.A. Nutritional needs in hot environments, nutritional needs in hot environments. 1993. [PubMed] [CrossRef]

Newton P.C.D., Lieffering M., Parsons A.J., Brock S.C., Theobald P.W., Hunt C.L., Luo D., Hovenden M.J. Selective grazing modifies previously anticipated responses of plant community composition to elevated CO2 in a temperate grassland. Global Change Biol. 2014;20:158–169. doi:10.1111/gcb.12301. [PubMed] [CrossRef] [Google Scholar]

Nielsen U.N., Stafford-Smith M., Metternicht G.I., Ash A., Baumber A., Boer M.M., Booth S., Burnside D., Churchill A.C., El Hassan M., Friedel M.H., Godde C.M., Kelly D., Kelly M., Leys J.F., McDonald S.E., Maru Y.T., Phelps D.G., Ridges M., Simpson G., Traill B., Walker B., Waters C.M., Whyte A.W. Challenges, solutions and research priorities for sustainable rangelands. Rangel. J. 2020:359–373. doi:10.1071/rj20059. [CrossRef] [Google Scholar]

Olsen J.R., Zepp L.J., Dager C.A. Climate impacts on inland navigation. World water congr. 2005 impacts glob. Clim. Chang. - Proc. 2005 World Water Environ. Resour. Congr. 2005;463 doi:10.1061/40792(173)463. [CrossRef] [Google Scholar]

Pan C.G., Kimball J.S., Munkhjargal M., Robinson N.P., Tijdeman E., Menzel L., Kirchner P.B. Role of surface melt and icing events in livestock mortality across Mongolia's semi-arid landscape. Rem. Sens. 2019;11:1–20. doi:10.3390/rs11202392. [CrossRef] [Google Scholar]

Parmesan C., Yohe G. A globally coherent fingerprint of climate change impacts across natural systems. Nature. 2003;421:37–42. doi:10.1038/nature01286. [PubMed] [CrossRef] [Google Scholar]

Patterson C.D., Guerin M.T. The effects of climate change on avian migratory patterns and the dispersal of commercial poultry diseases in Canada-Part i. Worlds. Poultry Sci. J. 2013;69:17–26. doi:10.1017/S0043933913000020. [CrossRef] [Google Scholar]

Patz J.A., Campbell-Lendrum D., Holloway T., Foley J.A. Impact of regional climate change on human health. Nature. 2005;438:310–317. doi:10.1038/nature04188. [PubMed] [CrossRef] [Google Scholar]

Peng S., Piao S., Shen Z., Ciais P., Sun Z., Chen S., Bacour C., Peylin P., Chen A. Precipitation amount, seasonality and frequency regulate carbon cycling of a semi-arid grassland ecosystem in Inner Mongolia, China: a modeling analysis. Agric. For. Meteorol. 178– 2013;179:46–55. doi:10.1016/j.agrformet.2013.02.002. [CrossRef] [Google Scholar]

Peñuelas J., Ciais P., Canadell J.G., Janssens I.A., Fernández-Martínez M., Carnicer J., Obersteiner M., Piao S., Vautard R., Sardans J. Shifting from a fertilization-dominated to a warming-dominated period. Nat. Ecol. Evol. 2017;1:1438–1445. doi:10.1038/s41559-017-0274-8. [PubMed] [CrossRef] [Google Scholar]

Polley H.W., Bailey D.W., Nowak R.S., Stafford-Smith M. Ecological consequences of climate change on rangelands. In: Briske D.D., editor. Rangeland Systems. Springer Series on Environmental Management; 2017. pp. 229–260. [CrossRef] [Google Scholar]

Polsky L., von Keyserlingk M.A.G. Invited review: effects of heat stress on dairy cattle welfare. J. Dairy Sci. 2017;100:8645–8657. doi:10.3168/jds.2017-12651. [PubMed] [CrossRef] [Google Scholar]

Prevéy J.S., Seastedt T.R. Seasonality of precipitation interacts with exotic species to alter composition and phenology of a semi-arid grassland. J. Ecol. 2014;102:1549–1561. doi:10.1111/1365-2745.12320. [CrossRef] [Google Scholar]

Proctor F., Lucchesi V. IIED/HIVOS; London, The Hague: 2012. Small-scale Farming and Youth in an Era of Rapid Rural Change. [Google Scholar]

Rashamol V.P., Sejian V., Pragna P., Lees A.M., Bagath M., Krishnan G., Gaughan J.B. Prediction models, assessment methodologies and biotechnological tools to quantify heat stress response in ruminant livestock. Int. J. Biometeorol. 2019;63:1265–1281. doi:10.1007/s00484-019-01735-9. [PubMed] [CrossRef] [Google Scholar]

Ray D.K., West P.C., Clark M., Gerber J.S., Prishchepov A.V., Chatterjee S. Climate change has likely already affected global food production. PloS One. 2019;14:1–18. doi:10.1371/journal.pone.0217148. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Reid R.S., Fernandez-Gimenez M.E., Galvin K.A. Dynamics and resilience of rangelands and pastoral peoples around the globe. Annu. Rev. Environ. Resour. 2014;39:217–242. doi:10.1146/annurev-environ-020713-163329. [CrossRef] [Google Scholar]

Ricciardi V., Wane A., Sidhu B.S., Goode C., Solomon D., McCullough E., Diekmann F., Porciello J., Jain M., Randall N., Mehrabi Z. A scoping review of research funding for small-scale farmers in water scarce regions. Nat. Sustain. 2020;3:836–844. doi:10.1038/s41893-020-00623-0. [CrossRef] [Google Scholar]

Ritzema R.S., Douxchamps S., Fraval S., Bolliger A., Hok L., Phengsavanh P., Long C.T.M., Hammond J., van Wijk M.T. Household-level drivers of dietary diversity in transitioning agricultural systems: evidence from the Greater Mekong Subregion. Agric. Syst. 2019;176:102657. doi:10.1016/j.agsy.2019.102657. [CrossRef] [Google Scholar]

Rivera-Ferre M.G., López-i-Gelats F., Howden M., Smith P., Morton J.F., Herrero M. Wiley Interdiscip. Rev. Clim. Chang; 2016. Re-framing the climate change debate in the livestock sector: mitigation and adaptation options. [CrossRef] [Google Scholar]

Robinson T.P., Thornton P.K., Franceschini G., Kruska R.L., Chiozza F., Notenbaert A., Cecchi G., Herrero M., Epprecht M., Fritz S., You L., Conchedda G., See L. 2011. Global Livestock Production Systems. Rome. [Google Scholar]

Rojas-Downing M.M., Nejadhashemi A.P., Harrigan T., Woznicki S.A. Climate change and livestock: impacts, adaptation, and mitigation. Clim. Risk Manag. 2017;16:145–163. doi:10.1016/j.crm.2017.02.001. [CrossRef] [Google Scholar]

Romeo A., Meerman J., Demeke M., Scognamillo A., Asfaw S. Linking farm diversification to household diet diversification: evidence from a sample of Kenyan ultra-poor farmers. Food Secur. 2016;8:1069–1085. doi:10.1007/s12571-016-0617-3. [CrossRef] [Google Scholar]

Romo-Barron C.B., Diaz D., Portillo-Loera J.J., Romo-Rubio J.A., Jimenez-Trejo F., Montero-Pardo A. Impact of heat stress on the reproductive performance and physiology of ewes: a systematic review and meta-analyses. Int. J. Biometeorol. 2019;63:949–962. doi:10.1007/s00484-019-01707-z. [PubMed] [CrossRef] [Google Scholar]

Rosenzweig C., Elliott J., Deryng D., Ruane A.C., Müller C., Arneth A., Boote K.J., Folberth C., Glotter M., Khabarov N., Neumann K., Piontek F., Pugh T.A.M., Schmid E., Stehfest E., Yang H., Jones J.W. Assessing agricultural risks of climate change in the 21st century in a global gridded crop model intercomparison. Proc. Natl. Acad. Sci. Unit. States Am. 2014;111:3268–3273. doi:10.1073/pnas.1222463110. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Rosenzweig C., Iglesias A., Yang X.B., Epstein P.R., Chivian E. Climate change and extreme weather events - implications for food production, plant diseases, and pests. Global Change Hum. Health. 2001;2:90–104. doi:10.1023/A:1015086831467. [CrossRef] [Google Scholar]

Saeed M., Abbas G., Alagawany M., Kamboh A.A., Abd El-Hack M.E., Khafa*ga A.F., Chao S. Heat stress management in poultry farms: a comprehensive overview. J. Therm. Biol. 2019;84:414–425. doi:10.1016/j.jtherbio.2019.07.025. [PubMed] [CrossRef] [Google Scholar]

Salih A.A.M., Baraibar M., Mwangi K.K., Artan G. Climate change and locust outbreak in East Africa. Nat. Clim. Change. 2020;10:584–585. doi:10.1038/s41558-020-0835-8. [CrossRef] [Google Scholar]

Salman R., Ferdinand T., Choularton R., Carter R. 2019. Transformative Adaptation in Livestock Production Systems. Washington, DC. [Google Scholar]

Sarhadian R. Small Grocery Store Integrated Energy Efficiency Improvements. Refrigeration and Thermal Test. Center, Southern California Edison; Irwindale, CA: 2004. [Google Scholar]

Sayre N.F., McAllister R.R., Bestelmeyer B.T., Moritz M., Turner M.D. Earth Stewardship of rangelands: coping with ecological, economic, and political marginality. Front. Ecol. Environ. 2013;11:348–354. doi:10.1890/120333. [CrossRef] [Google Scholar]

Schewe J., Gosling S.N., Reyer C., Zhao F., Ciais P., Elliott J., Francois L., Huber V., Lotze H.K., Seneviratne S.I., van Vliet M.T.H., Vautard R., Wada Y., Breuer L., Büchner M., Carozza D.A., Chang J., Coll M., Deryng D., de Wit A., Eddy T.D., Folberth C., Frieler K., Friend A.D., Gerten D., Gudmundsson L., Hanasaki N., Ito A., Khabarov N., Kim H., Lawrence P., Morfopoulos C., Müller C., Müller Schmied H., Orth R., Ostberg S., Pokhrel Y., Pugh T.A.M., Sakurai G., Satoh Y., Schmid E., Stacke T., Steenbeek J., Steinkamp J., Tang Q., Tian H., Tittensor D.P., Volkholz J., Wang X., Warszawski L. State-of-the-art global models underestimate impacts from climate extremes. Nat. Commun. 2019;10:1–14. doi:10.1038/s41467-019-08745-6. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Schweikert A., Chinowsky P., Espinet X., Tarbert M. Climate change and infrastructure impacts: comparing the impact on roads in ten countries through 2100. Procedia Eng. 2014;78:306–316. doi:10.1016/j.proeng.2014.07.072. [CrossRef] [Google Scholar]

Schweikert A., Chinowsky P., Kwiatkowski K., Espinet X. The infrastructure planning support system: analyzing the impact of climate change on road infrastructure and development. Transport Pol. 2014;35:146–153. doi:10.1016/j.tranpol.2014.05.019. [CrossRef] [Google Scholar]

Scoones I. Sustainable rural livelihoods: a framework for analysis. IDS Work. Pap. 1998;72:22. [Google Scholar]

Sejian V., Gaughan J., Baumgard L., Prasad C.S., editors. Climate Change Adaptation Livestock: Impact on and Mitigation. Springer India; New Delhi Heidelberg New York Dordrecht London: 2015. [CrossRef] [Google Scholar]

Selvaraj S., Ganeshamoorthi P., Pandiaraj T. Potential impacts of recent climate change on biological control agents in agro-ecosystem: a review. Int. J. Biodivers. Conserv. 2013;5:845–852. https://www.globalscienceresearchjournals.org/abstract/potential-impacts-of-recent-climate-change-on-biological-control-agents-in-agroecosystem-a-review-46423.html [Google Scholar]

Seré C., Steinfeld H. Animal Production and Health; Rome: 1996. World Livestock Production Systems: Current Status, Issues and Trends (No. 127)http://www.fao.org/3/a-w0027e.pdf [Google Scholar]

Sevi A., Caroprese M. Impact of heat stress on milk production, immunity and udder health in sheep: a critical review. Small Rumin. Res. 2012;107:1–7. doi:10.1016/j.smallrumres.2012.07.012. [CrossRef] [Google Scholar]

Sharma A., Tariq P.H., Kewalramani N., Kundu S.S. Livestock rearing on saline water. In: Dagar J.C., Sharma P.C., Sharma D.K., Singh A.K., editors. Innovative Saline Agriculture. Springer India; 2016. pp. 475–487. [CrossRef] [Google Scholar]

Sillmann J., Thorarinsdottir T., Keenlyside N., Schaller N., Alexander L.V., Hegerl G., Seneviratne S.I., Vautard R., Zhang X., Zwiers F.W. Understanding, modeling and predicting weather and climate extremes: challenges and opportunities. Weather Clim. Extrem. 2017;18:65–74. doi:10.1016/j.wace.2017.10.003. [CrossRef] [Google Scholar]

Sloat L.L., Gerber J.S., Samberg L.H., Smith W.K., Herrero M., Ferreira L.G., Godde C.M., West P.C. Increasing importance of precipitation variability on global livestock grazing lands. Nat. Clim. Change. 2018;8:214–218. doi:10.1038/s41558-018-0081-5. [CrossRef] [Google Scholar]

Smith K.R., Woodward A., Campbell-Lendrum D., Chadee D.D., Honda Y., Liu Q., Olwoch J.M., Revich B., Sauerborn R. Human health: impacts, adaptation, and co-benefit. In: Field C.B., Barros V.R., Dokken D.J., Mach K.J., Mastrandrea M.D., Bilir T.E., Chatterjee M., Ebi K.L., Estrada Y.O., Genova R.C., Girma B., Kissel E.S., Levy A.N., MacCracken S., Mastrandrea P.R., White L.L., editors. Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press; Cambridge, United Kingdom and New York, NY, USA: 2014. pp. 709–754.https://researchprofiles.canberra.edu.au/en/publications/human-health-impacts-adaptation-and-co-benefits [Google Scholar]

Smith M.R., Myers S.S. Impact of anthropogenic CO2 emissions on global human nutrition. Nat. Clim. Change. 2018;8:834–839. doi:10.1038/s41558-018-0253-3. [CrossRef] [Google Scholar]

Springmann M., Clark M., Mason-D’Croz D., Wiebe K., Bodirsky B.L., Lassaletta L., de Vries W., Vermeulen S.J., Herrero M., Carlson K.M., Jonell M., Troell M., DeClerck F., Gordon L.J., Zurayk R., Scarborough P., Rayner M., Loken B., Fanzo J., Godfray H.C.J., Tilman D., Rockström J., Willett W. Options for keeping the food system within environmental limits. Nature. 2018;562:519–525. doi:10.1038/s41586-018-0594-0. [PubMed] [CrossRef] [Google Scholar]

St-Pierre N.R., Cobanov B., Schnitkey G. Economic losses from heat stress by US livestock industries. J. Dairy Sci. 2003;86:E52–E77. doi:10.3168/jds.S0022-0302(03)74040-5. [CrossRef] [Google Scholar]

Stafford Smith M., Foran B. An approach to assessing the economic risk of different drought management tactics on a South Australian pastoral sheep station. Agric. Syst. 1992;39:83–105. doi:10.1016/0308-521X(92)90006-A. [CrossRef] [Google Scholar]

Suárez-Alemán A. Short sea shipping in today's Europe: a critical review of maritime transport policy Ancor. Marit. Econ. Logist. 2016;18:331–351. doi:10.1057/mel.2015.10. [CrossRef] [Google Scholar]

Tajul Baharuddin M.F., Taib S., Hashim R., Abidin M.H.Z., Rahman N.I. Assessment of seawater intrusion to the agricultural sustainability at the coastal area of Carey Island, Selangor, Malaysia. Arab. J. Geosci. 2013;6:3909–3928. doi:10.1007/s12517-012-0651-1. [CrossRef] [Google Scholar]

Tedeschi L.O., Fox D.G. vol. 1. XanEdu Publishing Inc; 2020. Ruminant nutrition system.https://www.amazon.com.au/Ruminant-Nutrition-System-Vol-Requirement/dp/1975077016 (An Applied Model for Predicting Nutrient Requirement and Feed Utilization in Ruminants). [Google Scholar]

The State of Victoria AgNote LC0077 - water supply for stock containment areas. 2018. [WWW Document]. URL http://agriculture.vic.gov.au/agriculture/farm-management/managing-dams/water-supply-for-stock-containment-areas.

The World Bank World development indicators. 2020. https://datacatalog.worldbank.org/dataset/world-development-indicators WWW Document]. URL.

Thomas D.S.G., Twyman C. Equity and justice in climate change adaptation amongst natural-resource-dependent societies. Global Environ. Change. 2005;15:115–124. doi:10.1016/j.gloenvcha.2004.10.001. [CrossRef] [Google Scholar]

Thornton P., Herrero M., Boone R. Altered grazing systems: pastoralism to conventional agriculture. In: Gibson D.J., Newman J.A., editors. Grasslands and Climate Change. Cambridge University Press; 2019. pp. 253–273. [Google Scholar]

Thornton P.K., Herrero M. Adapting to climate change in the mixed crop and livestock farming systems in sub-Saharan Africa. Nat. Clim. Change. 2015;5:830–836. doi:10.1038/nclimate2754. [CrossRef] [Google Scholar]

Thornton P.K., Herrero M. Climate change adaptation in mixed crop-livestock systems in developing countries. Glob. Food Sec. 2014;3:99–107. doi:10.1016/j.gfs.2014.02.002. [CrossRef] [Google Scholar]

Thornton P.K., Jones P.G., Ericksen P.J., Challinor A.J. Agriculture and food systems in sub-Saharan Africa in a 4°C+ world. Philos. Trans. A. Math. Phys. Eng. Sci. 2011;369:117–136. doi:10.1098/rsta.2010.0246. [PubMed] [CrossRef] [Google Scholar]

Thornton P.K., van de Steeg J., Notenbaert A., Herrero M. The impacts of climate change on livestock and livestock systems in developing countries: a review of what we know and what we need to know. Agric. Syst. 2009;101:113–127. doi:10.1016/j.agsy.2009.05.002. [CrossRef] [Google Scholar]

Toghiani S., Hay E.H., Roberts A., Rekaya R. Impact of cold stress on birth and weaning weight in a composite beef cattle breed. Livest. Sci. 2020;236:104053. doi:10.1016/j.livsci.2020.104053. [CrossRef] [Google Scholar]

Valente-Campos S., Spry D.J., Pascale Palhares J.C., Jakomin Rudez L.M., Umbuzeiro G. de A. Critical issues and alternatives for the establishment of chemical water quality criteria for livestock. Regul. Toxicol. Pharmacol. 2019;104:108–114. doi:10.1016/j.yrtph.2019.03.003. [PubMed] [CrossRef] [Google Scholar]

Valin H., Sands R.D., van der Mensbrugghe D., Nelson G.C., Ahammad H., Blanc E., Bodirsky B., Fujimori S., Hasegawa T., Havlik P., Heyhoe E., Kyle P., Mason-D’Croz D., Paltsev S., Rolinski S., Tabeau A., van Meijl H., von Lampe M., Willenbockel D. The future of food demand: understanding differences in global economic models. Agric. Econ. 2014;45:51–67. doi:10.1111/agec.12089. [CrossRef] [Google Scholar]

van der Spiegel M., van der Fels-Klerx H.J., Marvin H.J.P. Effects of climate change on food safety hazards in the dairy production chain. Food Res. Int. 2012;46:201–208. doi:10.1016/j.foodres.2011.12.011. [CrossRef] [Google Scholar]

von Wehrden H., Hanspach J., Ronnenberg K., Wesche K. Inter-annual rainfall variability in Central Asia - a contribution to the discussion on the importance of environmental stochasticity in drylands. J. Arid Environ. 2010;74:1212–1215. doi:10.1016/j.jaridenv.2010.03.011. [CrossRef] [Google Scholar]

Wagoner R.S., López-Gálvez N.I., de Zapien J.G., Griffin S.C., Canales R.A., Beamer P.I. An occupational heat stress and hydration assessment of agricultural workers in north Mexico. Int. J. Environ. Res. Publ. Health. 2020;17 doi:10.3390/ijerph17062102. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Walsh M.G., De Smalen A.W., Mor S.M. Climatic influence on anthrax suitability in warming northern latitudes. Sci. Rep. 2018;8:9269. doi:10.1038/s41598-018-27604-w. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Wang X., Stewart M.G., Nguyen M. Impact of climate change on corrosion and damage to concrete infrastructure in Australia. Climatic Change. 2012;110:941–957. doi:10.1007/s10584-011-0124-7. [CrossRef] [Google Scholar]

Ward D., McKague K. Water requirements of livestock. Ontario Minist. Agric. Food Rural Aff. 2019:8. http://www.omafra.gov.on.ca/english/engineer/facts/07-023.htm ISSN 1198-712X. [Google Scholar]

Wästerlund S. Forestry Working Paper; Rome: 2018. Managing Heat in Agricultural Work: Increasing Worker Safety and Productivity by Controlling Heat Exposure (No. 1)http://www.fao.org/forestry/harvesting/86024/en/ [Google Scholar]

Watson E.E., Kochore H.H., Dabasso B.H. Camels and climate resilience: adaptation in northern Kenya. Hum. Ecol. 2016;44:701–713. doi:10.1007/s10745-016-9858-1. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Weindl I., Lotze-Campen H., Popp A., Müller C., Havlík P., Herrero M., Schmitz C., Rolinski S., Havlik P., Herrero M., Schmitz C., Rolinski S. Livestock in a changing climate: production system transitions as an adaptation strategy for agriculture. Environ. Res. Lett. 2015;10 doi:10.1088/1748-9326/10/9/094021. [CrossRef] [Google Scholar]

WFP . 2019. Comprehensive Food Secruity and Vulnerability Analysis, Ethiopia, Addis Ababa, 2019. Addis Ababa, Ethiopia.https://reliefweb.int/report/ethiopia/ethiopia-comprehensive-food-security-and-vulnerability-analysis-cfsva-2019 [Google Scholar]

White B. Agriculture and the generation problem: rural youth, employment and the future of farming. IDS Bull. 2012;43:9–19. doi:10.1111/j.1759-5436.2012.00375.x. [CrossRef] [Google Scholar]

White W.B., McKeon G., Syktus J. Australian drought: the interference of multi-spectral global standing modes and travelling waves. Int. J. Climatol. 2003;23:631–662. doi:10.1002/joc.895. [CrossRef] [Google Scholar]

World Health Organization . 2014. Quantitative Risk Assessment of the Effects of Climate Change on Selected Causes of Death, 2030s and 2050s. [Google Scholar]

Wu X., Lu Y., Zhou S., Chen L., Xu B. Impact of climate change on human infectious diseases: empirical evidence and human adaptation. Environ. Int. 2016;86:14–23. doi:10.1016/j.envint.2015.09.007. [PubMed] [CrossRef] [Google Scholar]

Yengoh G.T., Ardö J. Climate change and the future heat stress challenges among smallholder farmers in East Africa. Atmosphere. 2020;11 doi:10.3390/atmos11070753. [CrossRef] [Google Scholar]

Young S.L. When an invasive plant fails to invade. Front. Ecol. Environ. 2015;13:450–451. doi:10.1890/1540-9295-13.8.450. [CrossRef] [Google Scholar]

Zeppel M.J.B., Wilks J.V., Lewis J.D. Impacts of extreme precipitation and seasonal changes in precipitation on plants. Biogeosciences. 2014;11:3083–3093. doi:10.5194/bg-11-3083-2014. [CrossRef] [Google Scholar]

Ziska L.H., Goins E.W. Elevated atmospheric carbon dioxide and weed populations in glyphosate treated soybean. Crop Sci. 2006;46:1354–1359. doi:10.2135/cropsci2005.10-0378. [CrossRef] [Google Scholar]

Ziska L.H., Pettis J.S., Edwards J., Hanco*ck J.E., Tomecek M.B., Clark A., Dukes J.S., Loladze I., Polley H.W. Rising atmospheric CO2 is reducing the protein concentration of a floral pollen source essential for north American bees. Proc. R. Soc. B Biol. Sci. 2016;283:20160414. doi:10.1098/rspb.2016.0414. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Ziska L.R., George K. Rising carbon dioxide and invasive, noxious plants: potential threats and consequences. World Resour. Rev. 2004;16:427–447. [Google Scholar]

Zvereva E.L., Kozlov M.V. Consequences of simultaneous elevation of carbon dioxide and temperature for plant-herbivore interactions: a metaanalysis. Global Change Biol. 2006;12:27–41. doi:10.1111/j.1365-2486.2005.01086.x. [CrossRef] [Google Scholar]

Queensland Government, 2019. The north-west Queensland Monsoon event of 26 January – 9 February 2019 : report of a landholder survey into impact and recovery.

Impacts of climate change on the livestock food supply chain; a review of the evidence (2024)
Top Articles
Set the default storage space for OneDrive users - SharePoint in Microsoft 365
Debt Collection Defense: Requiring That the Collector Document the Debt
Rpg Maker Fullscreen
Account Now Login In
Craigslist Farm And Garden Yakima Wa
Aldi Vs Costco: All Your Questions Answered
Ostedia
Lt4200 Huskee Manual
When His Eyes Opened Chapter 2694: Release Date, Spoilers & Where To Read? - OtakuKart
Power Outage Map Albany Ny
Living Room Furniture | Gavigan's Home Furnishings
Chowrastha - Indian Eatery Nashua Reviews
Cavalli Residential Flat Arabian Peninsula
Automation Personnel Services W2
Seo Glossary definition page
Steve Hytner Net Worth
O'reilly's Duquoin Illinois
Magicseaweed Jacksonville Fl
1818 West Taylor Street Chicago Il
1980 Monte Carlo For Sale Craigslist
Lthedom
Cayucos Craigslist
25Cc To Tbsp
043300109264
Pcc Skilled Nursing Login
Ramsey County Recordease
Nueces County Jail Inmate Search Vinelink
Craigslist Red Wing Mn
Doculivery Trinity Health
Security Awareness Training
Daryl Hannah Before and After Plastic Surgery: Face, Lips
Atlas Gradebook Uiuc
Vfr Town Of Salem
What Happened To Ed Hanna Wfmz
W.b. Crumel Funeral Home Obituaries
Megared Rewards
Amari Cooper Pfr
Bushnell Wingman Solid Orange Light
Bfads 2022 Walmart
Wyze Recover Deleted Events
Weather Past 3 Days
24Hrs Mcdonalds Near Me
45 Arch Street Akron Ohio
Oil Change Services | Jiffy Lube
Igumdrop Deepfake
Sarah Button Leaks
Www.gex-App-Ch
Rosalina Katrina Anderson
Td Bank Hours Weekend
Mugfaces Beaufort South Carolina
March 2023 Wincalendar
Worldfree4U Movies In
Latest Posts
Article information

Author: Allyn Kozey

Last Updated:

Views: 6604

Rating: 4.2 / 5 (43 voted)

Reviews: 90% of readers found this page helpful

Author information

Name: Allyn Kozey

Birthday: 1993-12-21

Address: Suite 454 40343 Larson Union, Port Melia, TX 16164

Phone: +2456904400762

Job: Investor Administrator

Hobby: Sketching, Puzzles, Pet, Mountaineering, Skydiving, Dowsing, Sports

Introduction: My name is Allyn Kozey, I am a outstanding, colorful, adventurous, encouraging, zealous, tender, helpful person who loves writing and wants to share my knowledge and understanding with you.